21.10.2014 Views

Physical Chemistry 3: — Chemical Kinetics — - Christian-Albrechts ...

Physical Chemistry 3: — Chemical Kinetics — - Christian-Albrechts ...

Physical Chemistry 3: — Chemical Kinetics — - Christian-Albrechts ...

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

<strong>Physical</strong> <strong>Chemistry</strong> 3:<br />

<strong>—</strong> <strong>Chemical</strong> <strong>Kinetics</strong> <strong>—</strong><br />

F. Temps<br />

Institute of <strong>Physical</strong> <strong>Chemistry</strong><br />

<strong>Christian</strong>-<strong>Albrechts</strong>-University Kiel<br />

Olshausenstr. 40<br />

D-24098 Kiel, Germany<br />

Email: temps@phc.uni.kiel.de<br />

URL: http://www.uni-kiel.de/phc/temps<br />

Summer Term 2013<br />

This scriptum contains notes for the lecture “<strong>Physical</strong> <strong>Chemistry</strong> 3:<br />

<strong>Chemical</strong> <strong>Kinetics</strong>” (MNF-chem0405) in the summer term 2013 at the<br />

Institute of <strong>Physical</strong> <strong>Chemistry</strong> of <strong>Christian</strong>-<strong>Albrechts</strong>-University (CAU)<br />

Kiel. This module is the last of the three semester lecture course<br />

in <strong>Physical</strong> <strong>Chemistry</strong> for B.Sc. students of <strong>Chemistry</strong> and Business<br />

<strong>Chemistry</strong> at CAU Kiel. It includes 3 hours weekly (3 SWS) for lectures<br />

and 1 hour weekly (1 SWS) for an exercise class.<br />

c° F. Temps (2004 <strong>—</strong> 2013) Last update: June 22, 2013


ii<br />

Contents<br />

Preface<br />

ix<br />

Disclaimer<br />

xi<br />

Organisational matters<br />

xii<br />

1 Introduction 1<br />

1.1 Types of chemical reactions 1<br />

1.2 Time scales for chemical reactions 2<br />

1.3 Historical events 5<br />

1.4 References 7<br />

2 Formal kinetics 8<br />

2.1 Definitions and conventions 8<br />

2.1.1 The rate of a chemical reaction 8<br />

2.1.2 Rate laws and rate coefficients 9<br />

2.1.3 The order of a reaction 10<br />

2.1.4 The molecularity of a reaction 11<br />

2.1.5 Mechanisms of complex reactions 12<br />

2.2 <strong>Kinetics</strong> of irreversible first-order reactions 16<br />

2.3 <strong>Kinetics</strong> of reversible first-order reactions (relaxation processes) 20<br />

2.4 <strong>Kinetics</strong> of second-order reactions 24<br />

2.4.1 A + A → products 24<br />

2.4.2 A + B → products 25<br />

2.4.3 Pseudo first-order reactions 26<br />

2.4.4 Comparison of first-order and second-order reactions 27<br />

2.5 <strong>Kinetics</strong> of third-order reactions 28<br />

2.6 <strong>Kinetics</strong> of simple composite reactions 29<br />

2.6.1 Consecutive first-order reactions 29<br />

2.6.2 Reactions with pre-equilibrium 32<br />

2.6.3 Parallel reactions 33<br />

2.6.4 Simultaneous first- and second-order reactions* 33<br />

2.6.5 Competing second-order reactions* 34


iii<br />

2.7 Temperature dependence of rate coefficients 35<br />

2.7.1 Arrhenius equation 35<br />

2.7.2 Deviations from Arrhenius behavior 37<br />

2.7.3 Examples for non-Arrhenius behavior 38<br />

2.8 References 40<br />

3 <strong>Kinetics</strong> of complex reaction systems 41<br />

3.1 Determination of the order of a reaction 41<br />

3.2 Application of the steady-state assumption 43<br />

3.2.1 The reaction H 2 + Br 2 → 2 HBr 43<br />

3.2.2 Further examples 45<br />

3.3 Microscopic reversibility and detailed balancing 47<br />

3.4 Generalized first-order kinetics* 48<br />

3.4.1 Matrix method* 48<br />

3.4.2 Laplace transforms* 53<br />

3.5 Numerical integration 57<br />

3.5.1 Taylor series expansions 58<br />

3.5.2 Euler method 58<br />

3.5.3 Runge-Kutta methods 59<br />

3.5.4 Implicit methods (Gear): 61<br />

3.5.5 Extrapolation methods (Bulirsch-Stoer) 62<br />

3.5.6 Example: Consecutive first-order reactions 62<br />

3.6 Oscillating reactions* 66<br />

3.6.1 Autocatalysis 66<br />

3.6.2 <strong>Chemical</strong> oscillations 67<br />

3.7 References 75<br />

4 Experimental methods 76<br />

4.1 End product analysis 76<br />

4.2 Modulation techniques 77<br />

4.3 The discharge flow (DF) technique 78<br />

4.4 Flash photolysis (FP) 82<br />

4.5 Shock tube studies of high temperature kinetics 85<br />

4.6 Stopped flow studies 86


iv<br />

4.7 NMR spectroscopy 87<br />

4.8 Relaxation methods 88<br />

4.9 Femtosecond spectroscopy 89<br />

4.10 References 93<br />

5 Collision theory 94<br />

5.1 Hard sphere collision theory 95<br />

5.1.1 The gas kinetic collision frequency 95<br />

5.1.2 Allowance for a threshold energy for reaction 99<br />

5.1.3 Allowance for steric hindrance 100<br />

5.2 Kinetic gas theory 101<br />

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

5.2.2 The rate of wall collisions in 3D 106<br />

5.3 Transport processes in gases 110<br />

5.3.1 The general transport equation 110<br />

5.3.2 Diffusion 112<br />

5.3.3 Heat conduction 113<br />

5.3.4 Viscosity 114<br />

5.4 Advanced collision theory 115<br />

5.4.1 Fundamental equation for reactive scattering in crossed molecular beams 116<br />

5.4.2 Crossed molecular beam experiments 117<br />

5.4.3 From differential cross sections () to reaction rate constants ( ) 120<br />

5.4.4 Laplace transforms in chemistry 121<br />

5.4.5 Bimolecular rate constants from collision theory 123<br />

5.4.6 Long-range intermolecular interactions 132<br />

6 Potential energy surfaces for chemical reactions 138<br />

6.1 Diatomic molecules 138<br />

6.2 Polyatomic molecules 139<br />

6.3 Trajectory Calculations 143<br />

7 Transition state theory 146<br />

7.1 Foundations of transition state theory 146<br />

7.1.1 Basic assumptions of TST 147


v<br />

7.1.2 Formal kinetic model 149<br />

7.1.3 The fundamental equation for ( ) 149<br />

7.1.4 Tolman’s interpretation of the Arrhenius activation energy 153<br />

7.2 Applications of transition state theory 154<br />

7.2.1 The molecular partition functions 154<br />

7.2.2 Example 1: Reactions of two atoms A + B → A···B ‡ → AB 156<br />

7.2.3 Example 2: The reaction F + H 2 → F···H···H ‡ → FH + H 157<br />

7.2.4 Example 3: Deviations from Arrhenius behavior (T dependence of ( ))* 159<br />

7.3 Thermodynamic interpretation of transition state theory 160<br />

7.3.1 Fundamental equation of thermodynamic transition state theory 160<br />

7.3.2 Applications of thermodynamic transition state theory 164<br />

7.3.3 Kinetic isotope effects 166<br />

7.3.4 Gibbs free enthalpy correlations 168<br />

7.3.5 Pressure dependence of 168<br />

7.4 Transition state spectroscopy 170<br />

7.5 References 174<br />

8 Unimolecular reactions 175<br />

8.1 Experimental observations 175<br />

8.2 Lindemann mechanism 178<br />

8.3 Generalized Lindemann-Hinshelwood mechanism 182<br />

8.3.1 Master equation 182<br />

8.3.2 Equilibrium state populations 184<br />

8.3.3 The density of vibrational states () 185<br />

8.3.4 Unimolecular reaction rate constant in the low pressure regime 188<br />

8.3.5 Unimolecular reaction rate constant in the high pressure regime 190<br />

8.4 The specific unimolecular reaction rate constants () 191<br />

8.4.1 Rice-Ramsperger-Kassel (RRK) theory 191<br />

8.4.2 Rice-Ramsperger-Kassel-Marcus (RRKM) theory 193<br />

8.4.3 The SACM and VRRKM model 198<br />

8.4.4 Experimental results 199<br />

8.5 Collisional energy transfer* 200<br />

8.6 Recombination reactions 201


vi<br />

8.7 References 202<br />

9 Explosions 203<br />

9.1 Chain reactions and chain explosions 203<br />

9.1.1 Chain reactions without branching 203<br />

9.1.2 Chain reactions with branching 204<br />

9.2 Heat explosions 207<br />

9.2.1 Semenov theory for “stirred reactors” ( =const) 207<br />

9.2.2 Frank<strong>—</strong>Kamenetskii theory for “unstirred reactors” 211<br />

9.2.3 Explosion limits 212<br />

9.2.4 Growth curves and catastrophy theory 212<br />

10 Catalysis 214<br />

10.1 <strong>Kinetics</strong> of enzyme catalyzed reactions (homogeneous catalysis) 214<br />

10.1.1 The Michaelis-Menten mechanism 215<br />

10.1.2 Inhibition of enzymes: 217<br />

10.2 <strong>Kinetics</strong> of heterogeneous reactions (surface reactions) 218<br />

10.2.1 Reaction steps in heterogeneous catalysis 218<br />

10.2.2 Physisorption and chemisorption 218<br />

10.2.3 The Langmuir adsorption isotherm 219<br />

10.2.4 Surface Diffusion 222<br />

10.2.5 Types of heterogeneous reactions 223<br />

11 Reactions in solution 227<br />

11.1 Qualitative model of liquid phase reactions 227<br />

11.2 Diffusion-controlled reactions 230<br />

11.2.1 Experimental observations 230<br />

11.2.2 Derivation of 230<br />

11.3 Activation controlled reactions 234<br />

11.4 Electron transfer reactions (Marcus theory) 235<br />

11.5 Reactions of ions in solutions 238<br />

11.5.1 Effect of the ionic strength of the solution 238<br />

11.5.2 Kinetic salt effect 238<br />

11.6 References 240


vii<br />

12 Photochemical reactions 241<br />

12.1 Basic radiative processes in a two-level system; Einstein coefficients 241<br />

12.2 Fluorescence quenching (Stern-Volmer equation) 242<br />

12.3 The radiative lifetime 0 and the fluorescence quantum yield Φ 243<br />

12.4 Radiationless processes in photoexcited molecules 244<br />

12.4.1 The Born-Oppenheimer approximation and its breakdown 244<br />

12.4.2 Avoided crossings of two potential curves of a diatomic molecule 248<br />

12.4.3 Quantum mechanical treatment of a coupled 2×2 system 250<br />

12.4.4 Radiationless processes in polyatomic molecules: Conical intersections 252<br />

13 Atmospheric chemistry* 256<br />

14 Combustion chemistry* 257<br />

15 Astrochemistry* 258<br />

16 Energy transfer processes* 259<br />

17 Reaction dynamics* 260<br />

18 Electrochemical kinetics* 261<br />

A Useful physicochemical constants 262<br />

B The Marquardt-Levenberg non-linear least-squares fitting algorithm 263<br />

C Solution of inhomogeneous differential equations 267<br />

C.1 General method 267<br />

C.2 Application to two consecutive first-order reactions 268<br />

C.3 Application to parallel and consecutive first-order reactions 270<br />

D Matrix methods 273<br />

D.1 Definition 273<br />

D.2 Special matrices and matrix operations 274<br />

D.3 Determinants 277<br />

D.4 Coordinate transformations 278<br />

D.5 Systems of linear equations 280<br />

D.6 Eigenvalue equations 282


viii<br />

E Laplace transforms 286<br />

F The Gamma function Γ () 288<br />

G Ergebnisse der statistischen Thermodynamik 290<br />

G.1 Boltzmann-Verteilung 291<br />

G.2 Mittelwerte 293<br />

G.3 Anwendung auf ein Zweiniveau-System 295<br />

G.4 Mikrokanonische und makrokanonische Ensembles 298<br />

G.5 Statistische Interpretation der thermodynamischen Zustandsgrößen 300<br />

G.6 Chemisches Gleichgewicht 301<br />

G.7 Zustandssumme für elektronische Zustände 303<br />

G.8 Zustandssumme für die Schwingungsbewegung 304<br />

G.9 Zustandssumme für die Rotationsbewegung 306<br />

G.10 Zustandssumme für die Translationsbewegung 307


ix<br />

Preface<br />

This scriptum contains lecture notes for the module “<strong>Physical</strong> <strong>Chemistry</strong> 3: <strong>Chemical</strong><br />

<strong>Kinetics</strong>” (chem0405), the last part of the 3 semester course in <strong>Physical</strong> <strong>Chemistry</strong> for<br />

B.Sc. students of <strong>Chemistry</strong> and Business <strong>Chemistry</strong> at CAU Kiel. It is assumed (but<br />

not formally required) that students have previously taken courses on chemical thermodynamics<br />

(“<strong>Physical</strong> <strong>Chemistry</strong> 1: <strong>Chemical</strong> Equilibrium” or equivalent) and elementary<br />

quantum mechanics (“<strong>Physical</strong> <strong>Chemistry</strong> 2: Structure of Matter” or equivalent) as<br />

well as “Mathematics for <strong>Chemistry</strong>” (parts 1 and 2, or equivalent). Module chem0405<br />

includes 3 hours weekly (3 SWS) for lectures and 1 SWS of exercises (56 contact hours<br />

combined). Students who started their studies at CAU Kiel should normally be in their<br />

4th semester.<br />

The scriptum gives a summary of the material covered in the scheduled lectures to<br />

allow students to repeat the material more economically. It covers basic material that<br />

all chemistry students should learn irrespective of their possible inclination towards inorganic,<br />

organic or physical chemistry, but goes well beyond the standard <strong>Physical</strong> <strong>Chemistry</strong><br />

textbooks used in the PC-1 and PC-2 courses. This is done in recognition of the<br />

established research focus at the Institute of <strong>Physical</strong> <strong>Chemistry</strong> at CAU to enable students<br />

to pursue their B.Sc. thesis project in <strong>Physical</strong> <strong>Chemistry</strong>. Some more specialized<br />

sections have been marked by asterisks and may be omitted on first reading towards the<br />

B.Sc. degree. Useful additional reference material is given in the Appendix.<br />

A brief lecture scriptum can never replace a textbook. For additional review, students<br />

are strongly encouraged (and at places required) to consult the recommended books<br />

on <strong>Chemical</strong> <strong>Kinetics</strong> beyond the standard <strong>Physical</strong> <strong>Chemistry</strong> textbooks. The book<br />

by Logan 1 is a comprehensive overview of the field of chemical kinetics and a nice<br />

introduction; it also has the advantage that it is written in German. The book by Pilling<br />

and Seakins 2 gives a good introduction especially to gas phase kinetics. The book by<br />

Houston 3 includes more information on the dynamics of chemical reactions. The book<br />

by Barrante 4 is highly recommended for everyone who wants to brush-up some math<br />

skills. A highly recommended web site for looking up mathematical definitions and<br />

recipies is the MathWorld online encyclopedia (http://mathworld.wolfram.com).<br />

The computer has become indispensable in modern research. This applies especially<br />

to <strong>Chemical</strong> <strong>Kinetics</strong>, where only computers can solve (by numerical integration) the<br />

coupled nonlinear partial differential equation systems that describe important complex<br />

reaction systems (e.g., related to combustion, the atmosphere or climate change). The<br />

integration of differential equations, a major task in chemical kinetics, can be performed<br />

using Wolfram Alpha (http://www.wolframalpha.com). More complicated problems can<br />

be solved with versatile computer algebra systems, e.g. (in order of increasing power)<br />

Derive, 5 MuPad, MathCad, 6 Maple, Mathematica 7 . A graphics program (e.g., Origin, 8<br />

1 S. R. Logan, Grundlagen der Chemischen Kinetik, Wiley-VCH, Weinheim, 1996.<br />

2 M. J. Pilling, P. W. Seakins, Reaction <strong>Kinetics</strong>, Oxford University Press, Oxford, 1995.<br />

3 P. L. Houston, <strong>Chemical</strong> <strong>Kinetics</strong> and Reaction Dynamics, McGraw-Hill, 2001.<br />

4 J. R. Barrante, Applied Mathematics for <strong>Physical</strong> <strong>Chemistry</strong>, Prentice Hall, 2003 ($ 47.11).<br />

5 Derive is available as shareware without charge.<br />

6 MathCad is installed in the PC lab.<br />

7 Mathematica has been used by the author for some of the lecture material.<br />

8 Student versions of Origin are available for ≈ 35 − 120 (depending on run time).


x<br />

or QtiPlot) may be needed for data evaluation and presentation. 9<br />

As practically all original scientific literature is published in English (British and American;<br />

in fact many chemists at some time in their professional lifes have to master the<br />

subtle differences between both), this scriptum is written in English to introduce students<br />

to the predominating language of science at an early stage. Students should also<br />

note that the working language of some of the research groups (including that of the<br />

author) in the <strong>Chemistry</strong> Department at CAU is English.<br />

Finally, I would like to encourage all students to ask questions when they appear, whether<br />

during the lecture or in the exercise classes!<br />

Kiel, June 22, 2013,<br />

F. Temps<br />

9 Excel also allows you to plot simple functions and data, but is not sufficient for serious scientific<br />

applications.


xi<br />

Disclaimer<br />

Use of this text is entirely at the reader’s risk. Some sections are incomplete, some<br />

equations still have to be checked, some figures need to be redrawn, references are still<br />

missing, etc. No warranties in any form whatever are given by the author.


xii<br />

Organisational matters<br />

• List of participants<br />

• Lecture schedule<br />

• Exams and grades<br />

<strong>—</strong> Homework assignments<br />

<strong>—</strong> Short questions<br />

<strong>—</strong> Written final exam (“Klausur”)<br />

I<br />

Table 1: List of participants.<br />

<strong>Chemistry</strong> (B.Sc.) 82<br />

Business <strong>Chemistry</strong> (B.Sc.) 25<br />

Biochemistry (B.Sc.)<br />

<strong>Chemistry</strong> (2x-B.Sc.)<br />

Physics (M.Sc.) 1<br />

Other ) 1<br />

Total 109<br />

) Not registered for exercise classes and final written exam (“Klausur”).<br />

I<br />

Figure 1: <strong>Physical</strong> chemistry courses at CAU Kiel.


xiii<br />

I<br />

Figure 2: Organisational matters.<br />

I<br />

Figure 3: Requirements and grading.


xiv<br />

I<br />

Figure 4: Lecture contents.<br />

I<br />

Figure 5: Table of contents (continued). a<br />

a Topics marked with asterisks will be omitted in this course for lack of time. They are the subject of<br />

specialized lectures for advanced students.


xv<br />

I<br />

Figure 6: Recommended textbooks on chemical kinetics.<br />

References to original publications and other recommended literature for further reading<br />

willbegivenattheendofeachchapter.


1<br />

1. Introduction<br />

1.1 Types of chemical reactions<br />

• Homogeneous reactions:<br />

<strong>—</strong> Reactions in the gas phase, e.g.,<br />

H 2 +Cl 2 → 2HCl<br />

<strong>—</strong> Reactions in the liquid phase, e.g., H + -transfer, reduction/oxidation reactions,<br />

hydrolysis, . . . (most of organic and inorganic chemistry)<br />

<strong>—</strong> Reactions in the solid phase (usually very slow, except for special systems<br />

and special conditions; not covered in this lecture).<br />

• Heterogeneous reactions:<br />

<strong>—</strong> Reactions at the gas-solid interface, e.g.,<br />

C()+12O 2<br />

() → CO()<br />

<strong>—</strong> Reactions at the gas-liquid interface, e.g., heterogeneous catalysis, atmospheric<br />

reactions on aerosols (e.g., Cl release into the gas phase from<br />

polar stratospheric clouds in antarctic spring)<br />

ClONO 2 ()+HCl() → Cl 2 ()+HNO 3 ()<br />

<strong>—</strong> Reactions at the liquid-solid boundary, e.g., solvation, heterogeneous catalysis,<br />

electrochemical reactions (electrode kinetics), . . .<br />

• Irreversible reactions, e.g.,<br />

H 2 +12O 2 → H 2 O<br />

• Reversible reactions, e.g.,<br />

H 2 +I 2 À 2HI<br />

• Composite reactions, e.g.,<br />

CH 4 +2O 2 → CO 2 +2H 2 O<br />

• Elementary reactions, e.g.,<br />

OH + H 2 → H 2 O+H<br />

CH 3 NC → CH 3 CN


1.2 Time scales for chemical reactions 2<br />

1.2 Time scales for chemical reactions<br />

A chemical reaction requires some “contact” between the respective reaction partners.<br />

The “time between contacts” thus determines the upper limit for the rate of a reaction.<br />

I<br />

Reactions in the gas phase:<br />

• Free mean pathlength À molecular dimension (diameter) .<br />

• Reaction rate is determined by time between collisions.<br />

• Time between two collisions @ :<br />

≈ 500 m s (1.1)<br />

y<br />

y<br />

≈ 10 −7 m (1.2)<br />

∆ ≈ ≈ 2 × 10 −10 s (1.3)<br />

• Gas phase reactions take place on the nanosecond time scale:<br />

I @ ≤ 1 ∆ =5× 109 s −1 (1.4)<br />

I<br />

Reactions in the liquid phase:<br />

• Molecules are permanently in contact with each other.<br />

• Typical distance between neighboring molecules (upper limit allowing for solvation<br />

shell):<br />

≈ 2 × 10 −8 10 −7 cm (1.5)<br />

• Maximal reaction rate is determined by the rate of diffusion of the reacting molecules<br />

in and out of the solvent cage.<br />

I<br />

Reactions in the solid phase:<br />

• Atoms and molecules localized on fixed lattice positions.<br />

• Reaction rate is determined by rate of diffusion (“hopping”) of the atoms and<br />

molecules via vacancies (unoccupied lattice positions, “Fehlstellen”) or interstitial<br />

sites (“Zwischengitterplätze”).<br />

• Hopping from one lattice position to another usually requires a corresponding high<br />

activation energy.<br />

• Solid state reactions are usually very slow.


1.2 Time scales for chemical reactions 3<br />

I<br />

Time scale of molecular vibrations:<br />

• The frequencies of the vibrational motions of the atoms in a molecule with respect<br />

to each other determine an upper limit for the rate of a unimolecular reaction.<br />

• Uncertainty principle (Fourier relation between frequency and time domain):<br />

y<br />

With = = ˜:<br />

∆ ∆ ≥ ~ (1.6)<br />

∆ ∆ ≥ 1<br />

2<br />

1<br />

∆ ≈<br />

2 ∆˜<br />

• Application to a C-C stretching vibration:<br />

y<br />

y<br />

(1.7)<br />

(1.8)<br />

˜(C-C) ≈ 1000 cm −1 (1.9)<br />

= = ˜ =3× 10 13 s −1 (1.10)<br />

= −1 =333 × 10 −14 s ≈ 33 fs (1.11)<br />

• Application to slow torsional motions of large molecular groups in solution:<br />

˜ ≈ 100 cm −1 (1.12)<br />

y<br />

∆ ≈ 330 fs (1.13)<br />

I Femtochemistry: We can observe ultrafast chemical reactions directly nowadays by<br />

following the motions of atoms or groups of atoms in the course of a reaction in real<br />

time by femtosecond spectroscopy and femtosecond diffraction techniques. This new<br />

area of “femtochemistry” has grown enormously in importance in the last decade, since<br />

the award of the Nobel prize in chemistry to A.H. Zewail in 1999.<br />

1 femtosecond =1fs=10 −15 s (1.14)<br />

I Timescale of e − -jump processes: Only electron jumps are faster than vibrational<br />

motions (by factor of roughly ).<br />

• Estimated time for − -jump to another orbital due to absorption/emission of a<br />

photon:<br />

∆ ≈ 10 −18 s (1.15)<br />

There are indications that electron correlation may slow this time to some<br />

10 −17 s. 10<br />

10 The group of F. Krausz (MPI for Quantum Optics, Munich) recently used attosecond metrology<br />

based on an interferometric method to measure a delay of 21 ± 5as in the photo-induced emission<br />

of electrons from the 2 orbitals of neon atoms with respect to emission from the 2 orbital by a<br />

100 eV light pulse (Schultze et al, Science 328, 1658 (2010)).


1.2 Time scales for chemical reactions 4<br />

• Other fast electron transfer processes:<br />

<strong>—</strong> Autoionization of a doubly excited state, by one electron falling to a lower<br />

orbital, while the other is excited into the ionization continuum.<br />

• But “chemical” intra- and intermolecular − -transfer reactions are again slower:<br />

<strong>—</strong> The electron jump itself is very fast (≈ 10 −18 s,asabove).<br />

<strong>—</strong> However, in the Franck-Condon limit, the reactions usually require substantial<br />

molecular rearrangements due to the structural differences between the<br />

reactants and products (⇒ Marcus theory, Section 11.4). The net reaction<br />

rate is thus again limited by the time scale of vibrational motion.<br />

I<br />

Conclusions:<br />

• The time scales of chemical reactions span the enormous range of some 30 orders<br />

of magnitude, from 1fs=10 −15 s up to 10 9 y=10 16 s!<br />

• In view of this huge span, a factor-to-two experimental uncertainty in an experimental<br />

measurement or a theoretical prediction of a rate constant really means<br />

rather little.<br />

• The enormous range of reaction rates, their complex dependencies on temperature,<br />

pressure, and species concentrations, and the question of the underlying<br />

reaction mechanisms that determine the reaction rates, make the field of chemical<br />

reaction kinetics and dynamics very attractive and rewarding!


1.3 Historical events 5<br />

1.3 Historical events<br />

I Sucrose inversion reaction (Wilhelmy, 1850): Kinetic investigation of the H + catalyzed<br />

hydrolysis of sucrose by monitoring the change as function of time of the optical<br />

activity of the solution from reactant (sucrose) to products (glucose and fructose). 11<br />

I<br />

Table 1.1: Optical activity parameters for the sucrose inversion reaction.<br />

C 12 H 22 O 11 +H 2 O −→ C 6 H 12 O 6 + C 6 H 12 O 6<br />

sucrose glucose fructose<br />

(1 M) ) +665 ◦ <strong>—</strong> +520 ◦ −920 ◦<br />

(1 M) ) Ø =+665 ◦ Ø = −200 ◦<br />

) (1 M) =specific angle of rotation of a 1 M solution at (Na-D).<br />

I<br />

Figure 1.1: Observed concentration-time profile of sucrose.<br />

11 This experiment is carried out in the <strong>Physical</strong> <strong>Chemistry</strong> Lab Course 1 at CAU (“Grundpraktikum”).


1.3 Historical events 6<br />

I<br />

Empirical rate law:<br />

ln <br />

0<br />

= − (1.16)<br />

Differentiation gives<br />

or<br />

y Empirical rate law:<br />

ln <br />

<br />

− ln 0<br />

= − ()=− (1.17)<br />

| {z } <br />

=0<br />

1 <br />

<br />

= − (1.18)<br />

<br />

= − (1.19)<br />

Reaction rate :<br />

≡− <br />

<br />

(1.20)<br />

Notice that even though there is one molecule of H 2 O taking part in the reaction, H2 O<br />

does not appear in the rate law, because it is practically constant in the dilute solution.<br />

I<br />

Figure 1.2: Milestones in chemical kinetics.


1.3 Historical events 7<br />

1.4 References<br />

Arrhenius 1889 S. A. Arrhenius, Über die Reaktionsgeschwindigkeit bei der Inversion<br />

von Rohrzucker durch Säuren, Z.Physik.Chem.4, 226 (1889).<br />

Bunker 1966 D. L. Bunker, Theory of Elementary Gas Reaction Rates, Pergamon,<br />

Oxford, 1966.<br />

Glasstone 1941 S. Glasstone, K. Laidler, H. Eyring, TheTheoryofRateProcesses,<br />

Mc Graw-Hill, New York, 1941.<br />

Homann 1975 K. H. Homann, Reaktionskinetik, Grundzüge der Physikalischen<br />

Chemie Bd. IV, R. Haase (Ed.), Steinkopf, Darmstadt, 1975.<br />

Houston 1996 P. L. Houston, <strong>Chemical</strong> <strong>Kinetics</strong> and Reaction Dynamics, McGraw-<br />

Hill, Boston, 2001.<br />

Johnston 1966 H. S. Johnston, Gas Phase Reaction Rate Theory, Ronald,NewYork,<br />

1966.<br />

Kassel 1932 L. S. Kassel, The <strong>Kinetics</strong> of Homogeneous Gas Reactions, Chem.Catalog,<br />

New York, 1932.<br />

Logan 1996 S. R. Logan, Grundlagen der Chemischen Kinetik, Wiley-VCH, Weinheim,<br />

1996.<br />

Pilling 1995 M. J. Pilling, P. W. Seakins, Reaction <strong>Kinetics</strong>, Oxford University Press,<br />

Oxford, 1995.<br />

Smith 1980 I. W. M. Smith, <strong>Kinetics</strong> and Dynamics of Elementary Gas Reactions,<br />

Butterworths, London, 1980.<br />

Steinfeld 1989 J. I. Steinfeld, J. S. Francisco, W. L. Haase, <strong>Chemical</strong> <strong>Kinetics</strong> and<br />

Dynamics, Prentice Hall, Englewood Cliffs, 1989.


1.4 References 8<br />

2. Formal kinetics<br />

2.1 Definitions and conventions<br />

2.1.1 The rate of a chemical reaction<br />

I<br />

Rate of a simple reaction of type A + B → C+D:<br />

≡− A<br />

= − B<br />

=+ C<br />

=+ D<br />

<br />

= − [A]<br />

<br />

= − [B]<br />

<br />

=+ [C]<br />

<br />

=+ [D]<br />

<br />

(2.1)<br />

(2.2)<br />

We will usually write the concentrations of a species A by using square brackets, i.e.,<br />

[A] and understand that [A] is time dependent, i.e., [A ()].<br />

I Definition of the rate of reaction: 12 Using the standard convention for the signs<br />

of the stoichiometric coefficients of the reactants (−1) and products (+1), we write a<br />

generic chemical reaction as<br />

1 B 1 + 2 B 2 + → B + +1 B +1 + (2.3)<br />

I<br />

The progress of the reaction from reactants to products can be described using the<br />

reaction progress coordinate :<br />

= <br />

(2.4)<br />

<br />

The time dependence leads to the definition of the reaction rate.<br />

Definition 2.1: The reaction rate is defined as<br />

≡ 1 <br />

<br />

<br />

(2.5)<br />

For the above reaction, therefore,<br />

= − 1 1<br />

| 1 | = − 1 2<br />

| 2 | = =+ 1 <br />

| | =+ 1 +1<br />

| +1 | <br />

= (2.6)<br />

I Units: 13<br />

• SI unit for the reaction rate:<br />

[] = mol m 3 s (2.7)<br />

12 Reaction rate = Reaktionsgeschwindigkeit. Von manchen Schulen wird auch der Begriff “Reaktionsrate”<br />

verwendet. Die bevorzugte “einzig wahre korrekte Übersetzung” wird von der jeweils anderen<br />

Seite sehr ernst genommen, allerdings gilt in Kiel in dieser Hinsicht Waffenstillstand.<br />

13 For general information on the SI system see (Mills 1988) or (Homann 1995).


2.1 Definitions and conventions 9<br />

• Convenient units for gas phase reactions:<br />

<strong>—</strong> molar units:<br />

[] = mol cm 3 s (2.8)<br />

<strong>—</strong> molecular units: 14 [] =molecules cm 3 s ⇒ cm −3 s −1 (2.9)<br />

<strong>—</strong> We prefer to use molar units for rate constants, because rate constants are<br />

inherently quantities that are “averages” of some other quantity, e.g., a molecular<br />

cross section weighted by the relevant statistical distribution function<br />

(usually the Boltzmann distribution). Rate constants are thus properties of<br />

an ensemble of molecules, they are not original molecular quantities.<br />

• Convenient units for liquid phase reactions:<br />

[] = mol ls=M s (2.10)<br />

2.1.2 Rate laws and rate coefficients<br />

I Therateequation(ratelaw): Experiments show that the reaction rate generally<br />

depends on the concentrations of the reactants. The functional dependence is expressed<br />

by the rate equation (also called rate law), 15 e.g.<br />

for reaction 2.3.<br />

= − 1 [B 1 ]<br />

= <br />

| 1 | <br />

(2.11)<br />

= ([B 1 ] [B 2 ] [B ] [B +1 ] ) (2.12)<br />

Theratelawsofcomplexreactions(i.e.,the ([B 1 ] , [B 2 ] , , [B ] , [B +1 ] , ))<br />

are often rather complicated; only rarely (as in the case of elementary reactions) are<br />

they simple and intuitive.<br />

Often (but not always), ([B 1 ] , [B 2 ] , , [B ] , [B +1 ] , ) can be written as a<br />

product of a rate coefficient and some power series of the concentrations, for<br />

example<br />

= − 1 [B 1 ]<br />

= [B 1 ] [B 2 ] ... (2.13)<br />

| 1 | <br />

Note that the exponents . . . are not usually equal to 1 2 ...<br />

I Therateconstant(ratecoefficient): The rate constant (also called rate coefficient)<br />

16 usually depends on temperature ( = ( )) and sometimes also on pressure<br />

( = ()), but not on the concentrations ( 6= ()).<br />

14 IUPAC has decided that “molecules” is not a unit; see (Mills 1989) or (Homann 1995).<br />

15 Rate law = Geschwindigkeits- oder Zeitgesetz.<br />

16 Rate constant = Geschwindigkeitskonstante. Es wird manchmal auch die Übersetzung “Ratenkonstante”<br />

verwendet (aber bitte nicht “Ratekonstante”).


2.1 Definitions and conventions 10<br />

I Rate laws for first-order and second-order reactions: Simple rate laws describe<br />

first-order and second-order reactions:<br />

• Example of a first-order reaction: CH 3 NC → CH 3 CN<br />

= − [CH 3NC]<br />

<br />

=+ [CH 3CN]<br />

<br />

= [CH 3 NC] (2.14)<br />

• Example of a second-order reaction: C 4 H 9 +C 4 H 9 → C 8 H 18<br />

= − 1 2<br />

[C 4 H 9 ]<br />

<br />

=+ [C 8H 18 ]<br />

<br />

= [C 4 H 9 ] 2 (2.15)<br />

I Question: How can we determine reaction rates and rate constants ?<br />

I<br />

Figure 2.1: The easiest <strong>—</strong> and always applicable <strong>—</strong> method for determining the reaction<br />

rate and rate coefficient is the tangent method. To determine , yousimplytake<br />

the slope of a plot of () at time ; then follows via Eq. 2.13.<br />

We’ll learn about better methods for determining for elementary reactions (first-order<br />

and second-order reactions) further below.<br />

2.1.3 The order of a reaction<br />

I<br />

Definition 2.2: The order of a reaction is defined as the sum of the exponents in the<br />

rate law (Eq. 2.13).


2.1 Definitions and conventions 11<br />

I Example: Considering a generic rate law, like<br />

− [A] = | | [A] [B] (2.16)<br />

<br />

the total order of the reaction is = + . We also say, the reaction is -th order in<br />

[A] and -th order in B, ...<br />

The order of a reaction may be determined by the tangent method (Fig. 2.1): Its effects<br />

on the reaction rate are observable by increasing (e.g., doubling) the concentrations of<br />

the reaction partners one by one. Better methods will be explored later (⇒ homework<br />

assignments).<br />

I<br />

Caveats:<br />

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

by an experimental measurement.<br />

• The order of a complex reaction cannot be determined by simply looking at the<br />

overall reaction.<br />

• There are reactions with integer order, like the first-order or second-order reactions<br />

above (Eqs. 2.14 and 2.15). However, the order can also be fractional, as for the<br />

following reactions:<br />

(1) CH 3 CHO → CH 4 +CO:<br />

[CH 4 ]<br />

<br />

= [CH 3 CHO] 32 (2.17)<br />

(2) H 2 +Br 2 → 2HBr:<br />

[HBr]<br />

<br />

= 0 [H 2 ][Br 2 ] 12<br />

1+ 00 [HBr] [Br 2 ]<br />

(2.18)<br />

• A negative order in a concentration means that the respective species acts as an<br />

inhibitor.<br />

• A positive order in a product concentration means that the reaction rate is enhanced<br />

by autocatalysis.<br />

2.1.4 The molecularity of a reaction<br />

I<br />

Definition 2.3: The molecularity of a reaction is the number of molecules in an elementary<br />

reaction step at the microscopic (molecular) level.


2.1 Definitions and conventions 12<br />

I<br />

There are only three possible cases for the molecularity (one example for each<br />

case):<br />

• monomolecular (= unimolecular) reactions:<br />

CH 3 NC → CH 3 CN (2.19)<br />

• bimolecular reactions:<br />

O+H 2 → OH + H (2.20)<br />

• termolecular reactions:<br />

H+O 2 +M→ HO 2 +M (2.21)<br />

• Events where more than three species come together in the right configuration<br />

and react are extremely unlikely and do not play a role.<br />

I Caveat: The order of a reaction can be integer or non-integer. However, the molecularity<br />

can be only 1 or 2, or at the most 3.<br />

2.1.5 Mechanisms of complex reactions<br />

A stoichiometric formula rarely describes the reactive events happening at the microscopic,<br />

molecular level. It is, for example, completely impossible that the combustion<br />

of 1 -heptane molecule with 11 molecules of O 2<br />

-C 7 H 16 +11O 2 → 7CO 2 +8H 2 O (2.22)<br />

in an internal combustion engine occurs in a single collision and gives 15 product molecules.<br />

Instead, the net reaction is the result of a complex sequence of a very large<br />

number (thousands, in this case) of unimolecular, bimolecular, and termolecular elementary<br />

reactions.<br />

The identification of the individual elementary reactions, the determination of their rates<br />

as a function of and , andtheverification of the ensuing reaction mechanism is one<br />

of the main tasks in the field of chemical kinetics.<br />

I<br />

Examples for complex reaction systems:<br />

• H 2 /O 2 system (Fig. 2.2),<br />

• Combustion of CH 4 (Fig. 2.3),<br />

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


2.1 Definitions and conventions 13<br />

I<br />

Figure 2.2: Reactions in the H 2 /O 2 system.<br />

I Rate equations for the H 2 /O 2 system: Assuming that we know the reaction mechanism<br />

(i.e., all elementary reactions contributing), we can write down the rate equations<br />

for all species. Doing so, we obtain a set of coupled non-linear differential equations<br />

(DE’s) which can be solved usually only numerically, if we want to describe the<br />

net reaction:<br />

[H]<br />

<br />

[OH]<br />

<br />

[O]<br />

<br />

= − 1 [H] [O 2 ]+ 2 [O] [H 2 ]+ 3 [OH] [H 2 ] − 4 [H] [O 2 ][M] (2.23)<br />

− 5 [H] [HO 2 ] − 6 [H] [HO 2 ]+<br />

= 1 [H] [O 2 ]+ 2 [O] [H 2 ] − 3 [OH] [H 2 ]+2 6 [H] [HO 2 ] (2.24)<br />

− 7 [OH] [HO 2 ] − 2 8 [OH] 2 [M]<br />

+ (2.25)<br />

= (2.26)


2.1 Definitions and conventions 14<br />

I Figure 2.3: Mechanism of CH 4 combustion (Warnatz 1996).<br />

I Figure 2.4: NO in combustion processes (Warnatz 1996).


2.1 Definitions and conventions 15<br />

I<br />

How we will proceed:<br />

• In the remainder of this Chapter, we shall learn how to analytically solve the rate<br />

equations for elementary reactions and for simple composite reactions.<br />

• In Chapter 3, we shall then consider different methods for solving coupled nonlinear<br />

differential equations for complex reaction systems.<br />

<strong>—</strong> These methods are well established for stationary reaction systems.<br />

<strong>—</strong> However, the modeling of instationary reaction systems (such as ignition<br />

processes, explosions and detonations, 3D or 4D models of atmospheric<br />

chemistry and climate evolution, . . . ) continues to be a tremendously challenging<br />

task.


2.2 <strong>Kinetics</strong> of irreversible first-order reactions 16<br />

2.2 <strong>Kinetics</strong> of irreversible first-order reactions<br />

A → B (2.27)<br />

I Solution of the rate equations for A and B:<br />

− [A]<br />

<br />

=+ [B]<br />

<br />

= [A] (2.28)<br />

Mass balance:<br />

[A]<br />

<br />

+ [B]<br />

<br />

=0 (2.29)<br />

Integration of the rate equation (Eq. 2.28):<br />

Z<br />

− [A] = (2.30)<br />

[A]<br />

Z<br />

Z<br />

ln [A] = −= − (2.31)<br />

y<br />

ln [A] = −+ (2.32)<br />

Initial value condition at =0:<br />

[A ( =0)]=[A] 0<br />

(2.33)<br />

y<br />

Solution for [A ()]:<br />

y<br />

Solution for [B ()]:<br />

y<br />

=ln[A] 0<br />

(2.34)<br />

ln [A] = −+ln[A] 0<br />

(2.35)<br />

[A ()] = [A] 0<br />

− (2.36)<br />

[B ()] = [A] 0<br />

− [A ()] (2.37)<br />

¡<br />

[B ()] = [A] ¢ 0 1 − <br />

−<br />

(2.38)


2.2 <strong>Kinetics</strong> of irreversible first-order reactions 17<br />

I<br />

Figure 2.5: <strong>Kinetics</strong> of first-order reactions.<br />

I Time constant: The rate coefficient has the dimension of an inverse time. Thus,<br />

= 1 <br />

(2.39)<br />

hasdimensionoftimeandisthetimeafterwhich[A] has dropped to 1 <br />

its initial value:<br />

[A] =<br />

=[A] 0<br />

− = [A] 0<br />

<br />

≈ [A] 0<br />

2718<br />

(=368 %) of<br />

(2.40)<br />

I Half lifetime: The half lifetime 12 is the time after which [A] has dropped to 1 2 of<br />

its initial value (Fig. 2.5):<br />

y<br />

£<br />

A<br />

¡<br />

= 12<br />

¢¤<br />

=[A]0 − 12<br />

= [A] 0<br />

2<br />

− 12<br />

= 1 2<br />

(2.41)<br />

(2.42)<br />

y<br />

12 =ln2 (2.43)<br />

12 = ln 2<br />

<br />

(2.44)


2.2 <strong>Kinetics</strong> of irreversible first-order reactions 18<br />

I<br />

Conclusion:<br />

Thehalflifetimeforafirst-order reactions is independent of the initial concentration!<br />

I Experimental determination of k:<br />

(1) Plot of ln [A] vs. gives a straight line with slope = − (see Fig. 2.5):<br />

ln ([A] [A] 0<br />

)=− (2.45)<br />

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

(2) Numerical fit of to the exponential decay curve using the Marquardt-Levenberg<br />

non-linear least-squares fitting algorithm (Fig. 2.6; see Appendix B for details).<br />

This method is useful in particular for analyzing experimental decay curves with<br />

a constant background term. Again, we do not need to know the absolute concentration<br />

of [A].<br />

Advantages: Can be applied in presence of a constant background signal (noise,<br />

weak DC signal). Can also be applied for analyzing data from pump-probe experiments,<br />

when the duration of the pump is not short to the decay (deconvolution).<br />

I Figure 2.6: Exponential decay curve of a particular vibration-rotation state of<br />

the CH 3 O radical resulting from its unimolecular dissociation reaction according to<br />

CH 3 O → H+H 2 CO (Dertinger 1995). The small box is the output box from a fit<br />

using the software package ORIGIN.


2.2 <strong>Kinetics</strong> of irreversible first-order reactions 19<br />

I<br />

Figure 2.7: Excited-state relaxation dynamics of the adenine DNA dinucleotide after<br />

UV excitation.<br />

I Note (and warning): Anybody who wants to do least-squares fitting is urged to<br />

consult appropriate literature before doing the fitting! The book by Bevington 17 is<br />

considered required reading for any physical chemistry graduate student.<br />

As the saying goes: With three parameters, you can draw an elephant. With a fourth<br />

parameter, you can make him walk.<br />

17 P. R. Bevington, D. K. Robinson, Data Reduction and Error Analysis for the <strong>Physical</strong> Sciences,<br />

McGraw-Hill, Boston, 1992.


2.3 <strong>Kinetics</strong> of reversible first-order reactions (relaxation processes) 20<br />

2.3 <strong>Kinetics</strong> of reversible first-order reactions (relaxation processes)<br />

A<br />

1 À<br />

−1<br />

B (2.46)<br />

I<br />

Rate equation:<br />

[A]<br />

<br />

= − 1 [A] + −1 [B] (2.47)<br />

I<br />

Equilibrium at t →∞:<br />

[A]<br />

<br />

y<br />

= − [B]<br />

<br />

= − 1 [A] + −1 [B] = 0 (2.48)<br />

[B] ∞<br />

[A] ∞<br />

= 1<br />

−1<br />

= (2.49)<br />

• Therateconstantforthereversereaction −1 can be calculated from 1 and .<br />

• Or the equilibrium constant can be determined from measurements of the<br />

forward and reverse rate constants. From ,wecanthendetermineimportant<br />

thermochemical quantities, for example for reactive free radicals, via<br />

= ( B ª )( C ª )<br />

=exp<br />

∙− ∆ ¸<br />

ª<br />

(2.50)<br />

( A ª )<br />

<br />

• Important note:<br />

<strong>—</strong> In the frequently encountered case that the number of species on the reactant<br />

and product side of a reaction differ, one has to be very careful with the<br />

units of and .<br />

Consider, for example, a gas phase reaction of the type<br />

In this case, has the dimension of mol cm 3 :<br />

B C<br />

<br />

= 1 [B] [C]<br />

= =<br />

−1 [A]<br />

0 has the dimension of bar:<br />

A → B+C (2.51)<br />

A<br />

<br />

= 1 B C<br />

A<br />

= 0 <br />

<br />

(2.52)<br />

0 = B C<br />

(2.53)<br />

A<br />

But the thermodynamic equilibrium constant is dimensionless:<br />

= ( ∙<br />

B ª )( C ª )<br />

= 0 <br />

( A ª ) =exp − ∆ ¸<br />

ª<br />

(2.54)<br />

ª <br />

<strong>—</strong> General relation between 0 ,and :<br />

= 0 × ¡ ª¢ −Σ ( ) = × ¡ ª¢ Σ ( )<br />

(2.55)


2.3 <strong>Kinetics</strong> of reversible first-order reactions (relaxation processes) 21<br />

I<br />

Solution of the rate equations for [A] and [B]:<br />

Returning to the time dependence of the reaction<br />

A<br />

1 À<br />

−1<br />

B (2.56)<br />

we now solve the rate equation<br />

[A]<br />

<br />

= − 1 [A] + −1 [B] (2.57)<br />

• Mass balance:<br />

[B] = [A] 0<br />

− [A] (2.58)<br />

• Integration:<br />

[A]<br />

<br />

= − 1 [A] + −1 [B] (2.59)<br />

= − 1 [A] + −1 ([A] 0<br />

− [A]) (2.60)<br />

= − ( 1 + −1 )[A]+ −1 [A] 0<br />

(2.61)<br />

y<br />

[A]<br />

( 1 + −1 )[A]− −1 [A] 0<br />

= − (2.62)<br />

• Solution by substitution:<br />

=( 1 + −1 )[A]− −1 [A] 0<br />

(2.63)<br />

y<br />

y<br />

<br />

[A] =( 1 + −1 ) (2.64)<br />

[A] =<br />

<br />

( 1 + −1 )<br />

(2.65)<br />

<br />

= −<br />

( 1 + −1 ) <br />

(2.66)<br />

<br />

= − ( 1 + −1 ) (2.67)<br />

ln = − ( 1 + −1 ) + (2.68)<br />

• Initial value condition at =0:<br />

y<br />

y<br />

= 0 (2.69)<br />

=ln 0 (2.70)<br />

<br />

0<br />

=exp[− ( 1 + −1 ) ] (2.71)


2.3 <strong>Kinetics</strong> of reversible first-order reactions (relaxation processes) 22<br />

I Solution for [A(t)] (Fig. 2.8):<br />

or<br />

( 1 + −1 )[A]− −1 [A] 0<br />

( 1 + −1 )[A] 0<br />

− −1 [A] 0<br />

=exp[− ( 1 + −1 ) ] (2.72)<br />

−1<br />

[A] − [A]<br />

1 + 0<br />

−1<br />

[A] 0<br />

−<br />

=exp[− ( 1 + −1 ) ] (2.73)<br />

−1<br />

[A]<br />

1 + 0<br />

−1<br />

This expression is not so easy to memorize, but we may recast it in a simple way:<br />

Using<br />

1<br />

= [B] ∞<br />

= [A] 0 − [A] ∞<br />

(2.74)<br />

−1 [A] ∞<br />

[A] ∞<br />

which gives<br />

[A] ∞<br />

=<br />

−1<br />

[A]<br />

1 + 0<br />

(2.75)<br />

−1<br />

we obtain<br />

[A ()] − [A] ∞<br />

=exp[− ( 1 + −1 ) ] (2.76)<br />

[A] 0<br />

− [A] ∞<br />

With ∆ [A] <br />

=[A()]−[A] ∞<br />

and ∆ [A] 0<br />

=[A] 0<br />

−[A] ∞<br />

, we write this result in compact<br />

form as<br />

∆ [A] <br />

∆ [A] 0<br />

=exp[− ( 1 + −1 ) ] (2.77)<br />

I<br />

Figure 2.8: <strong>Kinetics</strong> of reversible first-order reactions.


2.3 <strong>Kinetics</strong> of reversible first-order reactions (relaxation processes) 23<br />

I<br />

Relaxation processes:<br />

• The transition from a deviation from equilibrium into the equilibrium is called<br />

relaxation.<br />

• ( 1 + −1 ) −1 has the dimension of time and is called the relaxation time,<br />

=( 1 + −1 ) −1 (2.78)<br />

• The temporal evolution of relaxation processes from a nonequilibrium to the equilibtium<br />

state is determined by the sum of the respective forward and reverse rate<br />

constants!


2.4 <strong>Kinetics</strong> of second-order reactions 24<br />

2.4 <strong>Kinetics</strong> of second-order reactions<br />

2.4.1 A + A → products<br />

A+A→ P (2.79)<br />

I<br />

Solution of the rate equation:<br />

Integration:<br />

y<br />

Initial value condition at =0:<br />

− 1 [A]<br />

= [A] 2 (2.80)<br />

2 <br />

[A]<br />

2<br />

= −2 (2.81)<br />

[A]<br />

− 1 = −2 + (2.82)<br />

[A]<br />

[A ( =0)]=[A] 0<br />

(2.83)<br />

y<br />

Solution for [A ()] (Fig. 2.15):<br />

= − 1<br />

[A] 0<br />

(2.84)<br />

1<br />

[A] = 1 +2 (2.85)<br />

[A] 0<br />

I Experimental determination of k:<br />

(1) Graphical method:<br />

1<br />

[A] = 1<br />

[A] 0<br />

+2 (2.86)<br />

⇒ Plot of [A] −1 vs. gives a straight line with slope =2 (Fig. 2.9).<br />

⇒ We do need to know the absolute concentration of [A]!<br />

(2) We can again use a nonlinear least squares fitting method (Appendix B).


2.4 <strong>Kinetics</strong> of second-order reactions 25<br />

I<br />

Figure 2.9: <strong>Kinetics</strong> of second-order reactions.<br />

I<br />

Half lifetime:<br />

y<br />

y<br />

y<br />

£<br />

A<br />

¡<br />

12<br />

¢¤<br />

=<br />

[A] 0<br />

2<br />

(2.87)<br />

2<br />

[A] 0<br />

= 1<br />

[A] 0<br />

+2 12 (2.88)<br />

12 =<br />

1<br />

2 [A] 0<br />

(2.89)<br />

The half lifetime for a second-order reactions depends on the initial concentration!<br />

2.4.2 A + B → products<br />

A+B→ P (2.90)


2.4 <strong>Kinetics</strong> of second-order reactions 26<br />

I<br />

Solution of the rate equation:<br />

− [A]<br />

<br />

= − [B]<br />

<br />

= [A] [B] (2.91)<br />

(1) Substitution:<br />

=([A] 0<br />

− [A] <br />

)=([B] 0<br />

− [B] <br />

) (2.92)<br />

[A] = [A] 0<br />

− (2.93)<br />

y<br />

<br />

= ([A] 0 − )([B] 0<br />

− ) (2.94)<br />

(2) Integration using partial fractions (skipped here):<br />

(3) General solution:<br />

<br />

[B] 0<br />

[A] <br />

[A] 0<br />

[B] <br />

=exp([A] 0<br />

− [B] 0<br />

) (2.95)<br />

2.4.3 Pseudo first-order reactions<br />

In experimental studies of second-order reactions of the type A+B→ C, onelikesto<br />

use a high excess of one of the reactions partners (e.g., [B] À [A]) sothat[B] ≈ const<br />

Under this condition, the reaction is said to be pseudo first-order, because<br />

[B] = 0 (2.96)<br />

y Solution of [A ()]:<br />

[A] = [A] 0<br />

−0 =[A] 0<br />

−[B] (2.97)<br />

I<br />

Experimental investigation of pseudo first-order reactions:<br />

(1) Measurements of the pseudo first-order decay of [A] vs. at different concentrations<br />

of [B]: Plotofln [A] vs. give pseudo first-order rate constant 0 = [B]<br />

for each value of [B].<br />

(2) Plot of 0 = [B] vs. [B] gives straight line with slope .<br />

⇒ We do not need the absolute concentration of [A], only the concentration of [B] is<br />

needed.


2.4 <strong>Kinetics</strong> of second-order reactions 27<br />

I<br />

Figure 2.10: <strong>Kinetics</strong> of pseudo first-order reactions.<br />

2.4.4 Comparison of first-order and second-order reactions<br />

Consider a first-order and a second-order reaction with the same initial concentrations<br />

and the same initial rate at = 0. As shown in Fig. 2.11 below, the rate of the<br />

second-order reaction rapidly slows down as time increases.<br />

I<br />

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


2.5 <strong>Kinetics</strong> of third-order reactions 28<br />

2.5 <strong>Kinetics</strong> of third-order reactions<br />

y<br />

A+B+C→ D (2.98)<br />

− [A]<br />

<br />

= [A] [B] [C] (2.99)<br />

I Pseudo second-order combination reactions: In combination reactions of two<br />

species A + A → A 2 or A + B → AB (for example, the recombination of two free<br />

radicals), one of the species (bath gas M) is usually present in large excess ([M] À [A],<br />

[B]), so that the kinetics becomes pseudo second-order, giving the two cases:<br />

•<br />

•<br />

A+A+M→ A 2 +M (2.100)<br />

A+B+M→ AB + M (2.101)<br />

M plays the role of an inert collision partner which removes the excess energy from the<br />

newly formed A 2 or AB molecules.


2.6 <strong>Kinetics</strong> of simple composite reactions 29<br />

2.6 <strong>Kinetics</strong> of simple composite reactions<br />

2.6.1 Consecutive first-order reactions<br />

A −→ 1<br />

B (2.102)<br />

B −→ 2<br />

C (2.103)<br />

I<br />

Coupled differential equations:<br />

[A]<br />

<br />

= − 1 [A] (2.104)<br />

[B]<br />

<br />

=+ 1 [A] − 2 [B] (2.105)<br />

[C]<br />

<br />

=+ 2 [B] (2.106)<br />

In this case, since the DE’s are linear, there is an exact solution which we will examine<br />

first. We will also look at an approximate solution which can be obtained with the<br />

quasi steady-state approximation and at methods to obtain numerical solutions, which<br />

we have to use for complex non-linear inhomogeneous DE systems.<br />

I<br />

Mass balance:<br />

[A] + [B] + [C] = [A] 0<br />

(2.107)<br />

I<br />

Initial conditions:<br />

[A ( =0)] = [A] 0<br />

(2.108)<br />

[B ( =0)] = 0 (2.109)<br />

[C ( =0)] = 0 (2.110)<br />

I<br />

Exact solution:<br />

• First-order decay of [A]:<br />

y<br />

[A]<br />

<br />

= − 1 [A] (2.111)<br />

[A] = [A] 0<br />

− 1<br />

(2.112)<br />

• Inhomogeneous DE for [B]:<br />

[B]<br />

<br />

=+ 1 [A] − 2 [B] (2.113)


2.6 <strong>Kinetics</strong> of simple composite reactions 30<br />

• Solution for [B ()] 18 :<br />

⎧<br />

1 [A] ⎪⎨ 0<br />

¡<br />

<br />

− 1 − ¢ − 2 <br />

if 1 6= 2<br />

[B ()] =<br />

2 − 1<br />

(2.114)<br />

⎪⎩<br />

1 [A] 0<br />

− 1 <br />

if 1 = 2<br />

• Reaction products by mass balance:<br />

[C ()] = [A] 0<br />

− [A ()] − [B ()] = (2.115)<br />

I<br />

Figure 2.12: <strong>Kinetics</strong> of consecutive first-order reactions (case I).<br />

18 ThesolutionisdetailedinAppendixC.


2.6 <strong>Kinetics</strong> of simple composite reactions 31<br />

I<br />

Figure 2.13: <strong>Kinetics</strong> of consecutive first-order reactions (case II).<br />

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

[B] = 1 [A] 0<br />

2 − 1<br />

¡<br />

<br />

− 1 − − 2 ¢ , (2.116)<br />

One exponential describes the rise, the other the fall of [B]. It is sometimes implicitly,<br />

but wrongly assumed that the rise time corresponds to 1 and the decay time to 2 .<br />

The truth is that we cannot tell just from the shape of the concentration-time profile<br />

whether 1 or 2 correspond to the rise or to the fall of [B].<br />

I Quasi steady-state approximation: Under the condition that<br />

2 À 1 (2.117)<br />

we can find a simple solution for the DE’s. Under this condition, after a short initial<br />

induction time (see Fig. 2.13), the change of [B] is very small compared to that of<br />

[A]. Therefore,wehaveapproximately<br />

[B]<br />

<br />

≈ 0 (2.118)<br />

This important approximation is known as the (quasi)steady-state approximation. 19<br />

19 Deutsch: Quasistationaritätsannahme.


2.6 <strong>Kinetics</strong> of simple composite reactions 32<br />

I Quasi steady-state solution for [B (t)]: For we have<br />

y<br />

y<br />

[B]<br />

<br />

= 1 [A] − 2 [B] ≈ 0 (2.119)<br />

1 [A] ≈ 2 [B] (2.120)<br />

[B] <br />

= 1<br />

2<br />

[A] = 1<br />

2<br />

[A] 0<br />

− 1<br />

(2.121)<br />

For , [C ()] is therefore given by<br />

y<br />

[C]<br />

<br />

= 2 [B] ≈ 1 [A] (2.122)<br />

[C] = [A] 0<br />

¡<br />

1 − <br />

− 1 ¢ (2.123)<br />

2.6.2 Reactions with pre-equilibrium<br />

A<br />

1 À<br />

−1<br />

B (2.124)<br />

B 2 → C (2.125)<br />

I<br />

Equilibrium between [A] and [B]:<br />

1 −1 À 2 (2.126)<br />

y<br />

[B]<br />

[A] = 1<br />

= <br />

−1<br />

(2.127)<br />

[B] = [A] (2.128)<br />

I<br />

Time dependence of [C]:<br />

[C]<br />

<br />

= 2 [B] = 2 [A] (2.129)


2.6 <strong>Kinetics</strong> of simple composite reactions 33<br />

2.6.3 Parallel reactions<br />

A −→ 1<br />

B (2.130)<br />

A −→ 2<br />

C (2.131)<br />

A −→ 3<br />

D (2.132)<br />

y<br />

y<br />

and<br />

[A]<br />

<br />

= − ( 1 + 2 + 3 )[A] (2.133)<br />

[A] = [A] 0<br />

−( 1+ 2 + 3 )<br />

(2.134)<br />

[B] =<br />

[C] =<br />

[D] =<br />

1 [A] 0<br />

¡ ¢ 1 − <br />

−( 1 + 2 + 3 )<br />

( 1 + 2 + 3 )<br />

(2.135)<br />

2 [A] 0<br />

¡ ¢ 1 − <br />

−( 1 + 2 + 3 )<br />

( 1 + 2 + 3 )<br />

(2.136)<br />

3 [A] 0<br />

¡ ¢ 1 − <br />

−( 1 + 2 + 3 )<br />

( 1 + 2 + 3 )<br />

(2.137)<br />

Note that the decay of [A] is determined by the total removal rate constant =<br />

1 + 2 + 3<br />

2.6.4 Simultaneous first- and second-order reactions*<br />

A −→ 1<br />

C (2.138)<br />

A+A−→ 2<br />

D (2.139)<br />

y<br />

y<br />

[A]<br />

= − 1 [A] − 2 2 [A] 2<br />

<br />

(2.140)<br />

µ<br />

1<br />

[A] = −2 2 22<br />

+ + 1 <br />

exp (− 1 )<br />

1 1 [A] 0<br />

(2.141)<br />

I Approximation for k 1 t ¿ 1: Power series expansion of the exponential gives<br />

1<br />

[A] ≈ 1 µ <br />

1<br />

+ +2 2 × (2.142)<br />

[A] 0<br />

[A] 0


2.6 <strong>Kinetics</strong> of simple composite reactions 34<br />

2.6.5 Competing second-order reactions*<br />

A+A−→ 1<br />

A 2 (2.143)<br />

A+B−→ 2<br />

P 1 (2.144)<br />

B −→ 3<br />

P 2 (2.145)<br />

I<br />

Solution for the case [A] À [B]:<br />

(1)<br />

[A]<br />

<br />

= −2 1 [A] 2 − 2 [A] [B] (2.146)<br />

≈−2 1 [A] 2 (2.147)<br />

y<br />

or<br />

1<br />

[A] − 1<br />

[A] 0<br />

=2 1 (2.148)<br />

[A] = ¡ [A] 0 −1 +2 1 ¢ −1<br />

(2.149)<br />

=<br />

[A] 0<br />

1+2 1 [A] 0<br />

<br />

(2.150)<br />

(2)<br />

[B]<br />

= − 2 [A] [B] − 3 [B] (2.151)<br />

<br />

This is another inhomogeneous first-order DE, for which the solution can be found<br />

as outlined in Appendix C, giving<br />

[B] = [B] 0<br />

× (1 + 2 1 [A] 0<br />

) − 22 1<br />

× exp (− 3 ) (2.152)<br />

I<br />

Example:<br />

1<br />

CH 3 +CH 3 −→ C 2 H 6 (2.153)<br />

CH 3 +OH−→ 2<br />

CH 2 +H 2 O (2.154)<br />

OH −→ 3<br />

P 2 (2.155)<br />

In the experimental study of reaction 2.154 (Deters 1998), the rate coefficient for the<br />

reaction ( 2 according to Eq. 2.152) was fitted to measured OH concentration-time<br />

profiles in the presence of high concentrations of CH 3 using the Marquardt-Levenberg<br />

nonlinear least squares fitting algorithm (Press 1992). The rate coefficients 1 and<br />

3 were determined by independent measurements. The absolute CH 3 concentrations,<br />

which are needed for the evaluation of 2 according to Eq. 2.152, were obtained using<br />

a quantitative titration reaction.


2.7 Temperature dependence of rate coefficients 35<br />

2.7 Temperature dependence of rate coefficients<br />

2.7.1 Arrhenius equation<br />

I Arrhenius: 20 Svante Arrhenius (Arrhenius 1889) found the following empirical relation<br />

describing the temperature dependence of the reaction rate constant:<br />

ln <br />

(1 ) = − <br />

<br />

(2.156)<br />

I<br />

Definition 2.4: is called the Arrhenius activation energy of the reaction:<br />

= − ln <br />

(1 )<br />

(2.157)<br />

I<br />

It is very important to keep in mind the following points:<br />

• Eq. 2.156 is nothing else but a simple and very convenient way to describe the<br />

empirical temperature dependence of reaction rate constants!<br />

• Eq. 2.156 says that, within experimental error, a plot of ln vs. 1 should give<br />

a straight line. We will see later, however, that this is rarely exactly true. The<br />

study and the understanding of non-Arrhenius behavior is in fact an important<br />

topicinmodernkinetics.<br />

• as determined by Eqs. 2.156 and 2.157 is therefore a purely empirical quantity!<br />

• We will use Eq. 2.157 simply as the definition and recipy for determining from<br />

experimental data and to relate theoretical models to the experimental data!<br />

I The rationale behind the Arrhenius equation (2.156): The rationale for the<br />

Arrhenius equation is that molecules need additional energy so that chemical bonds can<br />

be broken, atoms can rearrange, and new bonds can be formed (Fig. 2.14).<br />

Van’t Hoff’s equation:<br />

∆ = 2 ln <br />

<br />

= 2 ln −→ <br />

<br />

= 2 ln −→ ←− <br />

<br />

− 2 ln ←− <br />

<br />

(2.158)<br />

(2.159)<br />

= −→ − ←− (2.160)<br />

Thus we can write<br />

ln <br />

<br />

= <br />

2 (2.161)<br />

20 Arrhenius is generally regarded as one of the founders of physical chemistry (together with Faraday,<br />

van’t Hoff, andOstwald).


2.7 Temperature dependence of rate coefficients 36<br />

or 21<br />

ln <br />

(1 ) = − <br />

<br />

(2.164)<br />

I<br />

Figure 2.14: Derivation of the Arrhenius equation.<br />

I Experimental determination of E (see Fig. 2.14): Aplotofln vs. 1 gives<br />

a straight line with slope − <br />

I Integration of Eq. 2.156: Assuming that is independent of ,Eq.2.156canbe<br />

integrated:<br />

ln = <br />

(2.165)<br />

<br />

2<br />

y<br />

ln = − <br />

+ (2.166)<br />

<br />

y<br />

( )= − <br />

(2.167)<br />

Equation 2.167 is used to represent ( ) over a limited -range using the two parameters<br />

and :<br />

• = Arrhenius activation energy (measure for the energy barrier of the reaction)<br />

• = pre-exponential factor (frequency factor, collision frequency).<br />

21<br />

y<br />

(1 )<br />

<br />

= − 1<br />

2 (2.162)<br />

= − 2 (1 ) (2.163)


2.7 Temperature dependence of rate coefficients 37<br />

I<br />

I<br />

It is important to realize, however, that the value of E is not equal to the<br />

true height of the potential energy barrier for the reaction!<br />

As said above, is an empirical quantity! Its relation to the true threshold energy 0 ,<br />

which is the minimum energy above which reaction may occur, will be a main subject<br />

in the theory of bimolecular and unimolecular reactions (Sections 5 <strong>—</strong> 8). There is also<br />

a statistical interpretation of (see Section 8).<br />

Figure 2.15: The reaction HCO → H + CO (G. Friedrichs).<br />

2.7.2 Deviations from Arrhenius behavior<br />

In general, is dependent on temperature, i.e., = ( ). For gas phase reactions,<br />

for example, there is at least the √ dependence of the hard sphere collision frequency.<br />

The -dependence of is often small compared to so that within experimental<br />

error the dependence of ( ) is determined by .<br />

If the -dependence of is not small compared to ,wewillhavetouseamore<br />

general expression for ( ), as explained in the following.<br />

I Generalized three-parameter expression for k (T ): To account for deviations<br />

from Arrhenius because of ( ), one conveniently employs the three parameter expression<br />

( )= − 0<br />

(2.168)<br />

with and 0 now being -independent quantities.<br />

Range of values for specific bimolecular reactions (justified by theory; see later):<br />

− 15 ≤ ≤ +3 (2.169)


2.7 Temperature dependence of rate coefficients 38<br />

I<br />

I<br />

Question: What is the relation of Eq. 2.168 with the Arrhenius equation?<br />

Answer: We have to evaluate the Arrhenius parameters from Eq. 2.168 according to<br />

the definitions of and :<br />

(1) :<br />

=+ 2 ln <br />

(2.170)<br />

<br />

=+ 2 µ<br />

ln + ln − <br />

0<br />

(2.171)<br />

<br />

<br />

y<br />

= 0 + (2.172)<br />

Onlyinthecasethat ¿ 0 can we neglect the term compared to 0 !<br />

(2) : From the expression for ,wehave<br />

y<br />

0 = − (2.173)<br />

( )= −( − )<br />

= − <br />

= − <br />

(2.174)<br />

(2.175)<br />

(2.176)<br />

Comparing coefficients, we find<br />

= (2.177)<br />

2.7.3 Examples for non-Arrhenius behavior<br />

Complex forming bimolecular reactions may show extreme non-Arrhenius behavior. An<br />

extreme case is the reaction OH + CO → H+CO 2 , which incidentally is the single<br />

most important reaction in hydrocarbon combustion next to H+O 2 → OH + O.


2.7 Temperature dependence of rate coefficients 39<br />

I Figure 2.16: The reaction OH + CO → H+CO 2 (Fulle 1996, Kudla 1992).<br />

Another case is encountered if tunneling plays a role (for example, H + transfer reactions<br />

at low temperatures). In this case, ( ) tends to a plateau value as → 0 (Eisenberger<br />

1991).


2.7 Temperature dependence of rate coefficients 40<br />

2.8 References<br />

Arrhenius 1889 S. A. Arrhenius, Über die Reaktionsgeschwindigkeit bei der Inversion<br />

von Rohrzucker durch Säuren, Z.Physik.Chem.4, 226 (1889).<br />

Bevington 1992 P. R. Bevington, D. K. Robinson, Data Reduction and Error Analysis<br />

for the <strong>Physical</strong> Sciences, McGraw-Hill, Boston, 1992.<br />

Dertinger 1995 S. Dertinger, A. Geers, J. Kappert, F. Temps, J. W. Wiebrecht,<br />

Rotation-Vibration State Resolved Unimolecular Dynamics of Highly Vibrationally<br />

Excited CH 3 O( 2 ): III. State Specific Dissociation Rates from Spectroscopic Line<br />

Profiles and Time Resolved Measurements, Faraday Discuss. Roy. Soc. 102, 31<br />

(1995).<br />

Deters 1998 R. Deters, M. Otting, H. Gg. Wagner, F. Temps, B. László, S. Dóbé,<br />

T. Bérces, A Direct Investigation of the Reaction CH 3 + OH at = 298K:<br />

Overall Rate Coefficient and CH 2 Formation, Ber. Bunsenges. Phys. Chem. 102,<br />

58 (1998).<br />

Eisenberger 1991 H. Eisenberger, B. Nickel, A. A. Ruth, W. Al-Soufi, K.H.Grellmann,<br />

M. Novo, Keto-Enol Tautomerization of 2-(2 0 -Hydroxyphenyl)benzoxazole<br />

and 2-(2 0 -Hydroxy-4 0 -methylphenyl)benzoxazole in the Triplet State: Hydrogen<br />

Tunneling and Isotope Effects. 2. Dual Phosphorescence Studies, J. Phys. Chem.<br />

95, 10509 (1991).<br />

Fulle 1996 D. Fulle, H. F. Hamann, H. Hippler, J. Troe, High Pressure Range of Addition<br />

Reactions of HO. II. Temperature and Pressure Dependence of the Reaction<br />

HO + CO À HOCO → H+CO 2 ,J.Chem.Phys.105, 983 (1996).<br />

Homann 1995 K. H. Homann, SI-Einheiten, VCH,Weinheim,1995.<br />

Kudla 1992 K. Kudla, A. G. Koures, L. B. Harding, G. C. Schatz, A quasiclassical<br />

trajectory study of OH rotational excitation in OH + CO collisions using ab initio<br />

potential surfaces, J. Chem. Phys. 96, 7465 (1992).<br />

Mills 1988 I. M. Mills et al., Units and Symbols in <strong>Physical</strong> <strong>Chemistry</strong>, Blackwell,<br />

Oxford, 1988.<br />

Press 1992 W. H. Press, S. A. Teukolsky, W. T. Vetterling, B. P. Flannery, Numerical<br />

Recipes in Fortran, Cambridge University Press, Cambridge, 1992. Versions are<br />

also available for C and Pascal.<br />

Warnatz 1996 J. Warnatz, U. Maas, R. W. Dibble, Combustion. <strong>Physical</strong> and <strong>Chemical</strong><br />

Fundamentals, Modeling and Simulation, Experiments, Pollutant Formation,<br />

Springer, Heidelberg, 1996.


2.8 References 41<br />

3. <strong>Kinetics</strong> of complex reaction systems<br />

Real reaction systems are usually much more complicated:<br />

• Net reaction mechanisms may consist of ≈ 100 - 10 000 elementary reactions.<br />

• No analytical solutions for the rate laws.<br />

• In favorable cases, one may be able to use certain approximations to simplify the<br />

reaction systems, like<br />

<strong>—</strong> apply steady-state approximation where possible,<br />

<strong>—</strong> take into account reactions in equilibrium,<br />

<strong>—</strong> take into account microscopic reversibility (detailed balancing).<br />

• For more acurate descriptions, however, one usually needs numerical solutions<br />

for the rate equations. Using modern computers, this is not a problem even for<br />

10 000 reactions, as long as the transport processes are not too complicated<br />

(as, for example, 1D reaction systems such premixed flames or 1D models of atmospheric<br />

chemistry). However, numerical solutions are still extremely challenging<br />

for instationary 3D reaction systems, such as<br />

<strong>—</strong> 3D models of atmospheric chemistry including daily/annual variations and<br />

couplings to the ocean,<br />

<strong>—</strong> ignition processes (internal combustion engines).<br />

3.1 Determination of the order of a reaction<br />

Any reaction mechanism which we may postulate has to explain the observed reaction<br />

order. Thus, the determination of the reaction order is a central topic.<br />

We explore some general methods for determining the reaction order by considering an<br />

example:<br />

= − [A] = [A] (3.1)<br />

<br />

I First-order and second-order reactions: Test whether plots of ln vs. or 1 vs.<br />

, give the straight lines expected for first-order or second-order reactions, respectively.<br />

I Log-log plot of reaction rate vs. concentration: We see from Eq. 3.1 that a<br />

log-log plot of vs. [A] gives a straight line with slope :<br />

lg ∝ lg [A] (3.2)


3.1 Determination of the order of a reaction 42<br />

I Half lifetime method (for n 6= 1): 22<br />

y<br />

Z<br />

− [A] = [A] (3.3)<br />

<br />

Z<br />

[A] − [A] = − (3.4)<br />

y<br />

−<br />

1<br />

( − 1) [A]−(−1) = −+ (3.5)<br />

Initial value condition at =0:<br />

[A ( =0)]=[A] 0<br />

(3.6)<br />

y<br />

y<br />

Half lifetime:<br />

y<br />

y<br />

y<br />

= − 1<br />

( − 1) [A]−(−1) 0<br />

(3.7)<br />

1<br />

[A] −1 − 1<br />

[A] −1<br />

0<br />

2 −1<br />

[A] −1<br />

0<br />

=( − 1) (3.8)<br />

£<br />

A<br />

¡<br />

12<br />

¢¤<br />

=<br />

1<br />

2 [A] 0<br />

(3.9)<br />

− 1<br />

[A] −1<br />

0<br />

12 = 2−1 − 1<br />

( − 1)<br />

=( − 1) 12 (3.10)<br />

1<br />

[A] −1<br />

0<br />

(3.11)<br />

lg 12 ∝−( − 1) lg [A] 0<br />

(3.12)<br />

y A log-log plot of 12 vs. [A] 0<br />

gives a straight line with slope −( − 1).<br />

22 We leave aside a usually occuring stoichiometric factor ν here, since we can always define νk = k 0<br />

as rate constant k 0 .


3.2 Application of the steady-state assumption 43<br />

3.2 Application of the steady-state assumption<br />

The steady state assumption allows us to reduce a large reaction mechanism to a smaller<br />

one by eliminating the respective intermediate species.<br />

3.2.1 The reaction H 2 + Br 2 → 2 HBr<br />

The reaction between H 2 and Br 2 (Bodenstein et al.; <strong>Christian</strong>sen, Polanyi & Herzfeld)<br />

runs as a thermal reaction or as a photochemically induced reaction (initiated by light).<br />

In the latter case, it is a chain reaction with a quantum yield of<br />

Φ ≈ 10 6 (3.13)<br />

I Definition 3.1: Quantum yield Φ:<br />

Φ =<br />

number of product molecules<br />

number of absorbed photons<br />

(3.14)<br />

I<br />

Reaction mechanism (elementary reactions) describing the H 2 -Br 2 system:<br />

Br 2 → Br + Br chain initiation by thermal dissociation (1)<br />

Br 2 + → Br + Br chain initiation by photolysis () (1 0 )<br />

Br + H 2 → HBr + H chain propagation (2)<br />

H+Br 2 → HBr + Br chain propagation (3)<br />

H+HBr → H 2 +Br inhibition (4)<br />

Br + HBr 6→ Br 2 +H endothermic (∆ ª =+170kJ mol) (6)<br />

y reaction (6) can be neglected<br />

Br + Br<br />

+M → Br 2 chain termination (5)<br />

(3.15)


3.2 Application of the steady-state assumption 44<br />

I Rates equations: With the steady state approximations (i.e., [X] ≈ 0) for[Br]<br />

and [H], weobtain 23<br />

[H]<br />

<br />

[Br]<br />

<br />

[HBr]<br />

<br />

= + 2 [Br] [H 2 ] − 3 [H] [Br 2 ] − 4 [H] [HBr] ≈ 0 (3.16)<br />

y 2 [Br] [H 2 ] − 4 [H] [HBr] = + 3 [H] [Br 2 ] (3.17)<br />

2 [Br] [H 2 ]<br />

y [H] <br />

=<br />

3 [Br 2 ]+ 4 [HBr]<br />

(3.18)<br />

(3.19)<br />

= +2 1 [Br 2 ] − 2 [Br] [H 2 ]+ 3 [H] [Br 2 ]+ 4 [H] [HBr]<br />

| {z }<br />

=−<br />

[H]<br />

≈0<br />

−2 5 [Br] 2 (3.20)<br />

= +2 1 [Br 2 ] − 2 5 [Br] 2 ≈ 0 (3.21)<br />

µ 12 1<br />

y [Br] <br />

= [Br 2 ]<br />

(3.22)<br />

5<br />

µ 12 1<br />

2 [H 2 ]<br />

y [H] <br />

= [Br 2 ] ×<br />

(3.23)<br />

5 3 [Br 2 ]+ 4 [HBr]<br />

= 2 [Br] [H 2 ]+ 3 [H] [Br 2 ] − 4 [H] [HBr] (3.24)<br />

Using 2 [Br] [H 2 ] − 4 [H] [HBr] = + 3 [H] [Br 2 ] (Eq. 3.17) and substituting [H] <br />

(Eq.<br />

3.18), the last line simplifies to<br />

[HBr]<br />

<br />

=2 3 [H] [Br 2 ] (3.25)<br />

µ 12 1<br />

2 [H 2 ]<br />

=2 3 [Br 2 ] × [Br 2 ] ×<br />

5 3 [Br 2 ]+ 4 [HBr]<br />

(3.26)<br />

= 2 2 ( 1 5 ) 12 [H 2 ][Br 2 ] 12<br />

1+( 4 3 )[HBr] [Br 2 ]<br />

(3.27)<br />

=<br />

0 [H 2 ][Br 2 ] 12<br />

1+ 00 [HBr] [Br 2 ]<br />

(3.28)<br />

I<br />

Result:<br />

with<br />

[HBr]<br />

<br />

= 0 [H 2 ][Br 2 ] 12<br />

1+ 00 [HBr] [Br 2 ]<br />

(3.29)<br />

0 =2 2 ( 1 5 ) 12 and 00 = 4 3 (3.30)<br />

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

23 For the photolysis induced reaction, 1 has to be replaced by 0 1 ,where is the light intensity and<br />

0 1<br />

the corresponding photolysis rate constants (related to the absorption coefficient).


3.2 Application of the steady-state assumption 45<br />

3.2.2 Further examples<br />

I<br />

I<br />

Exercise 3.1: The thermal decomposition of acetaldehyde according to the overall<br />

reaction<br />

CH 3 CHO → CH 4 +CO (3.31)<br />

is known to proceed via the following elementary steps (Rice-Herzfeld mechanism):<br />

CH 3 CHO → CH 3 +HCO (1)<br />

CH 3 +CH 3 CHO → CH 4 +CH 3 CO (2)<br />

(3.32)<br />

CH 3 CO → CH 3 +CO (3)<br />

CH 3 +CH 3 → C 2 H 6 (4)<br />

Derive an expression for the production rate of [CH 4 ]. (Hint: Use the steady-state<br />

approximations for [CH 3 CO] and [CH 3 ] and neglect the HCO for the derivation of the<br />

expression.) ¤<br />

A more complete mechanism takes into account the thermal decomposition of the HCO<br />

radicals according to HCO → H + CO (5) followed by the reaction H + CH 3 CHO →<br />

H 2 +CH 3 CO (6).<br />

Solution 3.1: With the steady state approximations for [CH 3 CO] and [CH 3 ],weobtain<br />

[CH 3 CO]<br />

<br />

= 2 [CH 3 ][CH 3 CHO] − 3 [CH 3 CO] ≈ 0 (3.33)<br />

y [CH 3 CO] <br />

= 2 [CH 3 ][CH 3 CHO]<br />

3<br />

(3.34)<br />

and<br />

y<br />

[CH 3 ]<br />

<br />

= 1 [CH 3 CHO] − 2 [CH 3 ][CH 3 CHO] (3.35)<br />

+ 3 [CH 3 CO] − 2 4 [CH 3 ] 2 (3.36)<br />

= 1 [CH 3 CHO] − 2 [CH 3 ][CH 3 CHO] (3.37)<br />

+ 3 2 [CH 3 ][CH 3 CHO]<br />

− 2 4 [CH 3 ] 2<br />

3<br />

(3.38)<br />

= 1 [CH 3 CHO] − 2 4 [CH 3 ] 2 ≈ 0 (3.39)<br />

µ 12 1 [CH 3 CHO]<br />

y [CH 3 ] <br />

=<br />

(3.40)<br />

2 4<br />

[CH 4 ]<br />

<br />

[CH 4 ]<br />

<br />

= 2 [CH 3 ][CH 3 CHO]<br />

= 2<br />

µ<br />

1<br />

2 4<br />

<br />

[CH 3 CHO] 32 (3.41)<br />

¥<br />

I Exercise 3.2: The thermal decomposition of ethane gives mainly H 2 +C 2 H 4 . In<br />

addition, small amounts of CH 4 are formed. The decay of the C 2 H 6 concentration can<br />

be described by the reaction steps<br />

C 2 H 6 → CH 3 +CH 3 (1)<br />

CH 3 +C 2 H 6 → CH 4 +C 2 H 5 (2)<br />

C 2 H 5 → C 2 H 4 +H (3)<br />

(3.42)<br />

H+C 2 H 6 → C 2 H 5 +H 2 (4)<br />

H+C 2 H 5 → C 2 H 6 (5)


3.2 Application of the steady-state assumption 46<br />

Using the steady-state approximations for the radicals [CH 3 ], [C 2 H 5 ],and[H], derivea<br />

rate law describing the overall decay of [C 2 H 6 ]. Note: From the C-H bond dissociation<br />

energies in C 2 H 6 and C 2 H 5 ,itisknownthat 3 À 1 . ¤<br />

I Solution 3.2:<br />

− [C 2H 6 ]<br />

<br />

=<br />

" µ<br />

3 <br />

2<br />

2 1 + 1<br />

4 + 12<br />

#<br />

1 3 4<br />

[C 2 H 6 ] (3.43)<br />

5<br />

¥


3.3 Microscopic reversibility and detailed balancing 47<br />

3.3 Microscopic reversibility and detailed balancing<br />

The principle of microscopic reversibility is a consequence of the time reversal<br />

symmetry of classical and quantum mechanics: If all atomic momenta of a molecule<br />

that has reached some product state from an initial state are reversed, the molecule will<br />

return to its initial state via thesamepath.<br />

A macroscopic manifestation of the principle of microscopic reversibility is detailed<br />

balancing which states that<br />

(1) in equilibrium, the forward and reverse rates of a process are equal, i.e., for a<br />

1<br />

reaction A 1 À A 2 ,<br />

−1<br />

and<br />

1 [A 1 ]= −1 [A 2 ] (3.44)<br />

(2) if molecules A 1 and A 2 are in equilibrium and A 2 and A 3 are in equilibrium, both<br />

with respect to each other, there is also equilibrium between A 1 and A 3 :<br />

A 1<br />

1<br />

À<br />

−1<br />

A 2<br />

2<br />

À<br />

−2<br />

A 3 (3.45)<br />

y<br />

A 1<br />

3<br />

À<br />

−3<br />

A 3 (3.46)<br />

Thus, we can write<br />

[A 2 ] <br />

= 1<br />

= 12<br />

[A 1 ] <br />

−1<br />

(3.47)<br />

[A 3 ] <br />

= 2<br />

= 23<br />

[A 2 ] <br />

−2<br />

(3.48)<br />

[A 3 ] <br />

= 3<br />

= 13<br />

[A 1 ] <br />

−3<br />

(3.49)<br />

= 12 23 = 1 2<br />

−1 −2<br />

(3.50)<br />

Detailed balancing allows us to simplify complex reaction systems by eliminating − 1<br />

out of a sequence of reversible reactions that are in equilibrium.


3.4 Generalized first-order kinetics* 48<br />

3.4 Generalized first-order kinetics*<br />

3.4.1 Matrix method*<br />

For coupled complex reaction systems with only first-order reactions, the solution of the<br />

rate equations can be reduced to an eigenvalue problem (Jost 1974). The solution of<br />

the rate equations, i.e., the calculation of the time-dependent concentration profiles,<br />

requires finding the eigenvalues and eigenvectors of the rate constant matrix of the<br />

system. 24<br />

Eigenvalue problems are readily solved in practice using numerical methods (Press 1992).<br />

We consider an example which we solved using conventional methods in Section 2.3.<br />

I<br />

Example:<br />

The rate equations<br />

A 1<br />

1<br />

À<br />

2<br />

A 2 (3.51)<br />

[A 1 ]<br />

<br />

= − 1 [A 1 ]+ 2 [A 2 ] (3.52)<br />

[A 2 ]<br />

<br />

=+ 1 [A 1 ] − 2 [A 2 ] (3.53)<br />

canbewritteninamorecompactformasamatrix equation:<br />

⎛ ⎞<br />

[A 1 ] µ µ <br />

⎜<br />

⎝ ⎟ −1 +<br />

[A 2 ]<br />

⎠ =<br />

2 [A1 ]<br />

(3.54)<br />

+ 1 − 2 [A 2 ]<br />

<br />

With the concentration vector<br />

and the rate constant matrix<br />

K =<br />

A =<br />

µ <br />

[A1 ]<br />

[A 2 ]<br />

µ <br />

−1 + 2<br />

+ 1 − 2<br />

(3.55)<br />

(3.56)<br />

this becomes<br />

·<br />

A = K · A (3.57)<br />

I<br />

Generalized matrix rate equation for coupled first-order reaction systems:<br />

·<br />

A = K · A (3.58)<br />

24 A brief overview on matrices and determinants is given in Appendix D.


3.4 Generalized first-order kinetics* 49<br />

I Ansatz: The ansatz for solving Eq. 3.58 assumes that we can express A in terms of<br />

another vector B, 25 A = P · B (3.59)<br />

where P · B means the dot product, using a rotation matrix P which is defined so that<br />

it diagonalizes K via the eigenvalue equation K·P = P · Λ , which by multiplication<br />

from the left with P −1 yields<br />

P −1 · K · P = Λ (3.60)<br />

The eigenvalue matrix Λ is a diagonal matrix of the form<br />

⎛ ⎞<br />

1 0 0<br />

Λ = ⎝ 0 2 0<br />

0 0 . ⎠ (3.61)<br />

..<br />

As defined by Eq. 3.60, its elements on the diagonal { 1 2 } are the eigenvalues of<br />

the rate constant matrix. ThematrixP is called the eigenvector matrix associated<br />

with K. The columns of P are the respective eigenvectors <br />

associated with the<br />

respective eigenvalues .<br />

I Solution: Inserting the ansatz<br />

A = P · B (3.62)<br />

into our matrix equation 26 ·<br />

A = K · A (3.63)<br />

we have<br />

or<br />

(P · B)<br />

<br />

Muliplication of this equation from the left by P −1 gives<br />

= K · P · B (3.64)<br />

P · ·<br />

B = K · P · B (3.65)<br />

P −1 · P · ·<br />

B = P −1 · K · P · B (3.66)<br />

or<br />

·<br />

B = Λ · B (3.67)<br />

This DE can be immediately integrated since Λ is diagonal. The solution is<br />

B = Λ B 0 (3.68)<br />

where Λ is a diagonal matrix with elements © 1 2 ª , and Λ B 0 means<br />

element-wise multiplication. The result gives the time dependence of B, i.e., B()<br />

starting from the initial values B 0 .<br />

25 The components of B will be seen to be linear combinations of the components of the concentration<br />

vector A. B is immediately obtained once we have the matrix of eigenvectors P (see below).<br />

26 The dot above a concentration vector indicates differentiation by .


3.4 Generalized first-order kinetics* 50<br />

Having B(), we can transform back into the original concentration basis by using<br />

B = P −1 · A (3.69)<br />

and<br />

B 0 = P −1 · A 0 (3.70)<br />

InsertingintooursolutionforB (Eq. 3.68), we obtain<br />

P −1 · A = Λ P −1 · A 0 (3.71)<br />

The final step is now a multiplication from the left with P which yields the solution for<br />

Eq. 3.58<br />

A = P · Λ P −1 · A 0 (3.72)<br />

I Note: This may look complicated to someone who is not used to it. However, once<br />

we have set up K, which is straightforward, everything is done by the computer: The<br />

computer computes the<br />

• eigenvalues of the rate constant matrix K,<br />

• the eigenvector matrix P oftherateconstantmatrixK,<br />

• all necessary inverse matrices, etc.,<br />

• and does all the matrix multiplications.<br />

All you usually have to do is to program the K matrix.<br />

I<br />

Application to our example (solution “by hand”):<br />

• Evaluation of the eigenvalue matrix Λ by solution of the eigenvalue equation<br />

det (K − Λ · I) =0 (3.73)<br />

with I as unity matrix:<br />

y<br />

¯ − 1 − + 2<br />

+ 1 − 2 − ¯ =0 (3.74)<br />

(− 1 − ) · (− 2 − ) − ( 1 · 2 )=0 (3.75)<br />

1 2 + 1 + 2 + 2 − 1 2 =0 (3.76)<br />

( +( 1 + 2 )) = 0 (3.77)<br />

y<br />

1 =0 (3.78)<br />

2 = − ( 1 + 2 ) (3.79)<br />

y<br />

Λ =<br />

µ <br />

0 0<br />

0 − ( 1 + 2 )<br />

(3.80)


3.4 Generalized first-order kinetics* 51<br />

• Conclusions:<br />

(1) The largest (negative) eigenvalue ( 2 in this case) determines the shortest<br />

time scale and thus the short-time evolution of the reactant concentrations!<br />

(2) The smallest (negative) eigenvalue ( 1 in this case) determines the longest<br />

time scale of the reaction, i.e., the evolution of the reactant concentrations<br />

at long times. The smallest (negative) eigenvalue thus determines the net<br />

reaction rate!<br />

(3) Note that in the example the concentrations at long times are constant<br />

(equilibrium!) and therefore 1 =0.<br />

• Evaluation of the eigenvectors p 1 und p 2 by inserting the eigenvalues { 1 2 }<br />

one by one into<br />

K · P = Λ · P (3.81)<br />

yields, for each and the respective p , a linear equation of the type<br />

K · p = · p (3.82)<br />

y<br />

y<br />

(K − ) · p =0 (3.83)<br />

(− 1 − ) 1 + 2 2 =0 (3.84)<br />

1 1 +(− 2 − ) 2 =0 (3.85)<br />

• Inserting 1 =0:<br />

− 1 1 + 2 2 =0 (3.86)<br />

1 1 + − 2 2 =0 (3.87)<br />

y<br />

y<br />

1 1 = 2 2 (3.88)<br />

1 = 2 · (3.89)<br />

2 = 1 · (3.90)<br />

y<br />

• Inserting 2 = − ( 1 + 2 ):<br />

p 1 =<br />

µ <br />

2 · <br />

1 · <br />

(3.91)<br />

(− 1 + 1 + 2 ) 1 + 2 2 =0 (3.92)<br />

1 1 +(− 2 + 1 + 2 ) 2 =0 (3.93)<br />

y<br />

2 1 + 2 2 =0 (3.94)<br />

1 1 + 1 2 =0 (3.95)


3.4 Generalized first-order kinetics* 52<br />

y<br />

y<br />

1 = − 2 (3.96)<br />

1 = (3.97)<br />

2 = − (3.98)<br />

y<br />

µ <br />

<br />

p 2 =<br />

−<br />

(3.99)<br />

• Eigenvector matrix (= rotation matrix P):<br />

µ <br />

2 <br />

P =<br />

1 −<br />

(3.100)<br />

• The arbitrary factor is determined by the initial conditions to be =1,sothat<br />

µ <br />

2 1<br />

P =<br />

(3.101)<br />

−1<br />

1<br />

• Solution for A ():<br />

A = P · Λ P −1 · A 0 (3.102)<br />

I<br />

Solution using symbolic algebra programs:<br />

• Rate constant matrix K:<br />

K =<br />

µ <br />

−1 + 2<br />

+ 1 − 2<br />

(3.103)<br />

• eigenvalues(K) ={− 1 − 2 0} :<br />

µ <br />

− (1 + <br />

Λ =<br />

2 )0<br />

0 0<br />

(3.104)<br />

• eigenvectors(K) =<br />

½µ<br />

−1<br />

1<br />

¾<br />

(Ã 1<br />

!)<br />

<br />

↔− 1 − 2 <br />

2<br />

1 ↔ 0:<br />

1<br />

⎛<br />

P = ⎝ −1 ⎞<br />

2<br />

1<br />

⎠ (3.105)<br />

+1 1<br />

• inverse(P) :<br />

⎛<br />

− ⎞<br />

1 2<br />

P −1 ⎜<br />

= ⎝<br />

1 + 2 1 + 2<br />

⎟<br />

1 1<br />

⎠ (3.106)<br />

1 + 2 1 + 2


3.4 Generalized first-order kinetics* 53<br />

• B 0 = P −1 A 0 :<br />

• Solution for B ():<br />

• Solution for A ():<br />

B 0 = P −1 A 0 = P −1 ·<br />

µ<br />

10<br />

0<br />

⎛<br />

<br />

⎜<br />

= ⎝<br />

− 10<br />

1<br />

1 + 2<br />

10<br />

1<br />

1 + 2<br />

⎞<br />

⎟<br />

⎠ (3.107)<br />

B = Λ B 0 (3.108)<br />

µ ⎛ ⎞<br />

1<br />

<br />

−( 1 + 2<br />

=<br />

) −<br />

0 ⎜ 10<br />

0 1 ⎝<br />

1 + 2<br />

⎟<br />

1<br />

⎠ (3.109)<br />

10<br />

1 + 2<br />

=<br />

⎛<br />

⎜<br />

⎝<br />

− 10<br />

1<br />

1 + 2<br />

× −( 1+ 2 )<br />

10<br />

1<br />

1 + 2<br />

× 1<br />

A = P · Λ P −1 · A 0 = P · Λ P −1 ·<br />

µ µ<br />

<br />

−( 1 + 2<br />

= P ·<br />

) 0<br />

P<br />

0 1<br />

−1 10<br />

·<br />

0<br />

= P ·<br />

= P ·<br />

=<br />

=<br />

⎜<br />

⎝<br />

⎞<br />

⎟<br />

⎠ (3.110)<br />

µ<br />

10<br />

<br />

µ ⎛ 1<br />

<br />

−( 1 + 2 ) −<br />

0 ⎜ 10<br />

0 1 ⎝<br />

1 + 2<br />

1<br />

10<br />

1 + <br />

⎛ µ<br />

⎞ 2<br />

−( 1+ 2 )<br />

1<br />

− 10<br />

1 + 2<br />

1<br />

µ<br />

10<br />

1<br />

1 + 2<br />

<br />

⎛<br />

2<br />

−( 1+ 2 )<br />

10 + 10 1<br />

⎜ 1 + 2 1 + 2<br />

⎝ 1<br />

−( 1+ 2 )<br />

10 − 10 1<br />

1 + 2<br />

⎛<br />

2 + 1 −( 1+ 2 )<br />

10<br />

⎜ 1 + 2<br />

⎝ 1 − −( 1+ 2 )<br />

10 1<br />

1 + 2<br />

1 + <br />

⎞ 2<br />

0<br />

⎞<br />

<br />

(3.111)<br />

(3.112)<br />

⎟<br />

⎠ (3.113)<br />

⎟<br />

⎠ (3.114)<br />

⎞<br />

⎟<br />

⎠ (3.115)<br />

⎟<br />

⎠ (3.116)<br />

This result is identical to the result for [A ()] obtained in Section 2.3.<br />

3.4.2 Laplace transforms*<br />

The concept of Laplace and inverse Laplace transforms is extremely useful in chemical<br />

kinetics because


3.4 Generalized first-order kinetics* 54<br />

(1) Laplace transforms provide a convenient method for solving differential equations,<br />

(2) Laplace transforms form the connection between microscopic molecular properties<br />

and statistically (Boltzmann) averaged quantities.<br />

I<br />

Definition 3.2: The Laplace transform L [ ()] of a function () is defined as the<br />

integral<br />

Z ∞<br />

() =L [ ()] = () − (3.117)<br />

where<br />

0<br />

• is a real variable,<br />

• () is a real function of the variable with the property () =0for 0,<br />

• is a complex variable,<br />

• () =L [ ()] is a function of the variable .<br />

I<br />

Definition 3.3: The inverse Laplace transform L −1 [ ()] of the function () is<br />

defined as the integral<br />

() =L −1 [ ()] = 1 Z<br />

2<br />

+∞<br />

−∞<br />

() (3.118)<br />

where<br />

• is an arbitrary real constant.<br />

I Relation between f (t) and F (p): Since L −1 [ ()] recovers the original function<br />

(), the pair of functions () and () is said to form a Laplace pair:<br />

L −1 [ ()] = L −1 [L [ ()]] = () (3.119)<br />

and<br />

L [ ()] = L £ L −1 [ ()] ¤ = () (3.120)<br />

A list of Laplace and inverse Laplace transforms of some functions can be found in<br />

Appendix E (Table E.1).


3.4 Generalized first-order kinetics* 55<br />

I Solution of differential equations using Laplace transforms: Consider the<br />

Laplace transform of the derivative () of a function (): 27<br />

L [() ] =<br />

Z ∞<br />

We can solve this equation for () via<br />

0<br />

()<br />

<br />

= () −¯¯∞<br />

− (3.121)<br />

Z<br />

0 − ∞<br />

= − ( =0)+<br />

0<br />

Z ∞<br />

0<br />

() ¡ −¢ (3.122)<br />

() − (3.123)<br />

= − 0 + () (3.124)<br />

() = L [() ]+ 0<br />

<br />

and recover the solution () by inverse Laplace transformation,<br />

() =L −1 ∙ L [() ]+0<br />

<br />

¸<br />

(3.125)<br />

(3.126)<br />

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

we consider the DE’s for two consecutive first-order reactions<br />

A 1<br />

→ B 2<br />

→ C (3.127)<br />

We already solved this problem previously (in Section 2.6.1) using conventional methods,<br />

and repeat the solution now using the Laplace transform method.<br />

Towards these ends, in this example, we shall denote the concentrations [A ()] and<br />

[B ()] as () and () and their Laplace transforms as () and (). Therespective<br />

initial concentrations shall be [A] 0<br />

= 0 and [B] 0<br />

= 0 =0. Thus we have:<br />

• DE for the concentration ():<br />

= 1 0 − 1 − 2 (3.128)<br />

• Laplace transform: () can be found using Eq. 3.124:<br />

L [] =− 0 + () (3.129)<br />

Taking the Laplace transform of the r.h.s. of Eq. 3.128: 28<br />

L [] =L £ 1 0 − 1 ¤ − L [ 2 ()] (3.130)<br />

= 1 0 L £ − 1 ¤ − 2 L [ ()] = 1 0<br />

+ 1<br />

− 2 () (3.131)<br />

= − 0 + () (Eq. 3.129) (3.132)<br />

27 We used in line 2: integration by parts; in line 3: the differivative ( − ) = − − ;inline4:<br />

the definition () =L [ ()] and the abbreviation ( =0)= 0 .<br />

28 We used L [ ]= 1 from Table E.1 and L [ ()] = ().<br />


3.4 Generalized first-order kinetics* 56<br />

Since we have 0 =0from the initial condition, this becomes<br />

()+ 2 () = 1 0<br />

+ 1<br />

(3.133)<br />

• Solution for ():<br />

() = 1 0<br />

1<br />

+ 1<br />

1<br />

+ 2<br />

(3.134)<br />

• Inverse Laplace transform: 29<br />

() =L [ ()] = 1 0<br />

2 − 1<br />

¡<br />

<br />

− 1 − − 2 ¢ (3.135)<br />

which is the same as obtained in Section 2.6.1.<br />

∙<br />

¸<br />

29 FromTableE.1,weseethatL −1 1<br />

1 ¡<br />

= − ¢ .<br />

( − )( − ) ( − )


3.5 Numerical integration 57<br />

3.5 Numerical integration<br />

I Initial value problems: For complex reaction systems, we have to integrate the coupled<br />

nonlinear DE systems numerically using the computer. We start with the known<br />

initial values of the concentrations at time 0 and the associated DE’s ( · ) and want<br />

to compute the values of at time 0 . In numerics, these problems are known as<br />

initial value problems. With additional conditions, for example for the spatial distributions<br />

of the concentration (fixed values at ( 0 0 0 1 )), they become boundary<br />

value problems.<br />

In solving these problems, we generally have to deal with systems of stiff nonlinear<br />

coupled (ordinary) differential equations (DEsystemsaresaidtobestiff, ifthe<br />

rate constants (more precisely: the eigenvalues) vary by many orders of magnitude;<br />

under these conditons, the changes of the dependable variables differ by many orders of<br />

magnitude).<br />

Thus we have to solve<br />

• equations (kinetic equations + transport processes), with<br />

• variables (concentrations), and the given<br />

• initial values and boundary values.<br />

I<br />

Applications of numerical integration methods:<br />

• Combustion chemistry:<br />

<strong>—</strong> Stationary combustion (for example diffusion flames),<br />

<strong>—</strong> Instationary combustion (ignition processes, turbulent combustion) - first<br />

successful attempts have been made,<br />

<strong>—</strong> Problem: For all but the simplest fuels, we have to deal with hundreds of<br />

species and thousands of elementary reactions.<br />

• Atmospheric chemistry:<br />

<strong>—</strong> 2D “box” models ( ( )) are straightforward.<br />

<strong>—</strong> 4D models are required (time, latitude, longitude, height)!<br />

<strong>—</strong> Huge problems, when it comes to<br />

∗ feedback with oceans, etc.?<br />

∗ do we know all reactions?<br />

∗ heterogeneous reactions?<br />

• <strong>Physical</strong> organic chemistry:<br />

<strong>—</strong> Elucidation of reaction mechanisms.<br />

• Macromolecular chemistry (including industrial macromolecular synthesis):<br />

<strong>—</strong> Prediction of chain length and other propertiesofpolymersandco-polymers.


3.5 Numerical integration 58<br />

I<br />

Survey of methods of numerical integration methods:<br />

• explicit methods (e.g., Runge-Kutta),<br />

• implicit methods (e.g., Gear),<br />

• extrapolation methods (e.g., Richardson, Bulirsch-Stoer, Deuflhard).<br />

In the following, we consider the basics of the most convenient methods.<br />

3.5.1 Taylor series expansions<br />

The Taylor expansion is the basis for all numerical integration methods. We start from<br />

the initial value of ( 0 ) at some time 0<br />

( 0 )= 0 (3.136)<br />

and want to determine the value ( 0 + ) at time 0 + . UsingtheDEfor <br />

we expand in a Taylor series around 0 :<br />

Recursion formula:<br />

·<br />

() =( ) (3.137)<br />

( 0 + ) = ( 0 )+ · · ( 0 0 )+ 2<br />

2! · ··<br />

( 0 0 )+ (3.138)<br />

= ( 0 )+ · ( 0 0 )+ 2<br />

2! · ·<br />

( 0 0 )+ (3.139)<br />

+1 = + · ( )+ 2<br />

2! · ·<br />

( )+ (3.140)<br />

The first-order term ( ) is given by the rate equation at , the second-order<br />

term ( ·<br />

) (which is taken into account for example by the 2 order Runge-Kutta<br />

method) and higher order terms (⇒ higher order Runge-Kutta methods) have to be<br />

evaluated numerically by finite differences.<br />

3.5.2 Euler method<br />

The Euler method is based on a 1 order Taylor expansion. If the step size is small<br />

enough, the higher order terms can be neglected:<br />

( 0 + ) = ( 0 )+ · ·<br />

( 0 )+O( 2 ) (3.141)<br />

Initial value problem:<br />

·<br />

= ( ) (3.142)


3.5 Numerical integration 59<br />

with<br />

( 0 )= 0 (3.143)<br />

Result by substituting ·<br />

( 0 )= ( 0 ):<br />

( 0 + ) = ( 0 )+( 0 )+O ¡ 2¢ (3.144)<br />

Recursion formula (see Fig. 3.1):<br />

+1 = + (3.145)<br />

+1 = + ( ) (3.146)<br />

Problem: Large errors for practical step sizes.<br />

I<br />

Figure 3.1: Euler method.<br />

3.5.3 Runge-Kutta methods<br />

I<br />

2 order Runge-Kutta method:<br />

Initial value problem:<br />

with<br />

·<br />

= ( ) (3.147)<br />

( 0 )= 0 (3.148)


3.5 Numerical integration 60<br />

2 order Runge-Kutta ansatz = Taylor expansion to 2 order:<br />

( 0 + ) = ( 0 )+ · ( 0 )+ 2 ··<br />

<br />

2! ( 0 )+O( 3 ) (3.149)<br />

Ã<br />

≈ ( 0 )+ ·<br />

·<br />

( 0 )+ 2 ( 0 + ) − ·<br />

!<br />

( 0 )<br />

+ O( 3 ) (3.150)<br />

2 <br />

= ( 0 )+ ³ ·(0 + )+ ·( 0 )´<br />

2 | {z }<br />

=2× slope of () at midpoint<br />

+ O( 3 ) (3.151)<br />

Recursion formula:<br />

+1 = + (3.152)<br />

+1 = + 2 + O( 3 ) (3.153)<br />

1 = ( ) (3.154)<br />

2<br />

µ<br />

= <br />

+ 2 + 1<br />

2<br />

<br />

(3.155)<br />

I<br />

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


3.5 Numerical integration 61<br />

I 4 order Runge-Kutta method: In practice, one frequently uses the 4 order<br />

Runge-Kutta method.<br />

I<br />

Figure 3.3: Illustration of the 4 order Runge-Kutta method.<br />

3.5.4 Implicit methods (Gear):<br />

The Runge-Kutta method becomes unstable for stiff DE systems. In this case, it is<br />

preferable to employ implicit (also called iterative or backward) integration methods.<br />

The starting point is the implicit form of the Euler method,<br />

( 0 + ) = ( 0 )+ · ·<br />

( 0 + )+O( 2 ) (3.156)<br />

For the initial value problem ·<br />

= ( ) with ( 0 )= 0 this leads to the recursion<br />

formula:<br />

+1 = + (3.157)<br />

+1 = + ( +1 +1 ) (3.158)<br />

Using the so-called predictor-corrector methods, +1 is predicted in a first step by<br />

starting from using one of the explicit methods described above. The obtained trial<br />

value (0)<br />

+1 is then corrected by using a backward integration scheme (Gear 1971a, Gear<br />

1971a, Press 1992) to obtain a better estimate +1. (1) Then, and (1)<br />

+1 are used for<br />

a new prediction-corrrection cycle, giving +1. (2) The procedure is repeated until the<br />

difference between successive iterations falls below some tolerance limit .<br />

The predictor-corrector method of Gear has been the method of choice for a long time<br />

(Gear 1971a, Gear 1971a).


3.5 Numerical integration 62<br />

3.5.5 Extrapolation methods (Bulirsch-Stoer)<br />

The idea behind extrapolation methods is that the final answer for +1 is approximated<br />

by an analytical function in the step size (Richardson extrapolation). The best strategy<br />

is to use a rational extrapolating function as the ansatz, and not a polynomial function<br />

(Bulirsch-Stoer). This function is determined and fitted to the problem of interest by<br />

performing calculations with various values of . The extrapolating function is then<br />

used for extrapolating from to +1 .<br />

The optimal strategy for stiff DE systems is to follow the Bulirsch-Stoer method (Stoer<br />

1980) in the form implemented by Deuflhard (Bader 1983, Deuflhard 1983, Deuflhard<br />

1985). For Fortran subroutines, see (Press 1992).<br />

I<br />

Figure 3.4: The principle of extrapolation methods.<br />

3.5.6 Example: Consecutive first-order reactions<br />

As an example, we solve the DE’s for two consecutive first-order reactions<br />

A 1<br />

−→ B 2<br />

−→ C (3.159)<br />

(a)exactly,usingsymbolicalgebrasoftwarelikeMathCad,Maple,Mathematica,or<br />

MuPad (the math engine built-in into SWP, the program used to write this script) and<br />

(b) using numeric integration.


3.5 Numerical integration 63<br />

I<br />

Exact solution using symbolic algebra software:<br />

(1) To set-up the initial value problem for the program, 30 we enter the following a<br />

1 × 6 matrix:<br />

⎡<br />

0 ⎤<br />

= − 1 <br />

0 =+ 1 − 2 <br />

0 =+ 2 <br />

(3.160)<br />

⎢ (0) = 0 ⎥<br />

⎣<br />

(0) = 0<br />

⎦<br />

(0) = 0<br />

(2) To obtain the solution, we select the menu item "SolveODE, Exact". The program<br />

gives the output:<br />

Exact solution is:<br />

⎡<br />

⎢ () = 1 µ<br />

⎤<br />

− 1<br />

0 1 − 0 1 2 + 0 2 − 2<br />

1<br />

1 2 − 1 2 − ⎥ 1<br />

⎢<br />

⎣<br />

− 1<br />

− 2<br />

() = 0 1 − 0 1<br />

2 − 1 2 − 1<br />

() = 1 µ<br />

<br />

− 1<br />

0 1 2 − 0 2 − 1<br />

1<br />

1 2 − 1 2 − 1<br />

⎥<br />

⎦<br />

(3.161)<br />

(3) After simplifying, factorizing, and rewriting the output, we obtain<br />

⎡<br />

() = 0<br />

¡ ¢ ⎤<br />

2 − 1 + 1 − 2<br />

− 2 − 1<br />

2 − 1 ⎢<br />

1<br />

¡ ¢<br />

⎣ () = 0 <br />

− 2<br />

− − 1 ⎥<br />

1 − ⎦<br />

2<br />

() = 0 − 1<br />

(3.162)<br />

We immediately recognize these results from section 2.<br />

I Numerical solutions: Numerical integration is performed by the different symbolic<br />

algebra programs using one of the methods outlined in the preceding subsection. We<br />

have to specify the DE’s, the initial values, and the values for the rate constants. 31<br />

(1) We adopt the rate constant values 1 =100 and 2 =100.<br />

(2) We define the problem (now using capital letters):<br />

⎡<br />

0 ⎤<br />

= −<br />

0 =+ − 10<br />

0 =+10<br />

⎢ (0) = 1 ⎥<br />

⎣<br />

(0) = 0<br />

⎦<br />

(0) = 0<br />

(3.163)<br />

30 The different software packages differinthewaysinwhichtheproblemshavetobesetup.<br />

31 SWP can handle the problem only if the rate constant values are explicitly entered in the DE’s.<br />

Other software does better.


3.5 Numerical integration 64<br />

, Functions defined: <br />

y 1.0<br />

0.5<br />

0.0<br />

0 2 4<br />

(3) Application of "SolveODE, Numeric" gives the output:<br />

, Functions defined: <br />

This means that the program now “knows” these functions.<br />

(4) We can plot , , and using the Plot2D menu point:<br />

1<br />

c<br />

0.75<br />

0.5<br />

0.25<br />

0<br />

0<br />

1.25<br />

2.5<br />

3.75<br />

5<br />

t<br />

(Weaddtheplotsfor and by dragging both symbols to the plot window)<br />

(5) With 0 =1, 1 =1,and 2 =10, we also plot the exact solutions , , and<br />

= 0<br />

2 − 1<br />

¡<br />

2 − 1 + 1 − 2<br />

− 2 − 1 ¢ (3.164)<br />

1<br />

¡<br />

= 0 <br />

− 2<br />

− ¢ − 1<br />

1 − 2<br />

(3.165)<br />

= 0 − 1<br />

(3.166)


3.5 Numerical integration 65<br />

1<br />

c<br />

0.75<br />

0.5<br />

0.25<br />

0<br />

0<br />

1.25<br />

2.5<br />

3.75<br />

5<br />

t<br />

(6) The rest is just some fixing up: We combine the two plots, with an increased<br />

number of points plotted, add axis labels and colors:<br />

1<br />

c(t)<br />

0.75<br />

0.5<br />

0.25<br />

0<br />

0<br />

1.25<br />

Numerical solution of the DE system defined by Eq. 3.163. Full lines: numeric<br />

solutions. Dotted lines: exact solutions.<br />

2.5<br />

t<br />

3.75<br />

5<br />

The exact and the numerical solutions are identical within numeric precision.


3.6 Oscillating reactions* 66<br />

3.6 Oscillating reactions*<br />

The reactant and product concentrations normally approach equilibrium in a smooth<br />

(often exponenential) fashion. Under special circumstances, however, the concentrations<br />

may also show oscillations and/or bistability as they approach equilibrium. Such<br />

behavior can occur only if there is a chemical feedback by autocatalysis.<br />

3.6.1 Autocatalysis<br />

Example for autocatalysis:<br />

A+B → B+B (3.167)<br />

Rate law:<br />

− [A]<br />

<br />

= [A] [B] (3.168)<br />

To solve this DE, we substitute for [B ()] using the mass balance relation [A] 0<br />

−[A ()] =<br />

[B ()] − [B] 0<br />

and introduce =[A()] as reaction progress variable:<br />

[B ()] = [A] 0<br />

+[B] 0<br />

− [A ()] (3.169)<br />

=[A] 0<br />

+[B] 0<br />

− (3.170)<br />

This gives the DE<br />

y<br />

Z<br />

− <br />

= ([A] 0 +[B] 0<br />

− ) (3.171)<br />

Z<br />

<br />

([A] 0<br />

+[B] 0<br />

) − = − (3.172)<br />

2<br />

Solution:<br />

or<br />

[B ()] =<br />

µ <br />

1 [A]0 [B ()]<br />

ln<br />

= − (3.173)<br />

[A] 0<br />

+[B] 0<br />

[B] 0<br />

[A ()]<br />

[A] 0<br />

+[B] 0<br />

1+([A] 0<br />

[B] 0<br />

)exp[− ([A] 0<br />

+[B] 0<br />

) ]<br />

(3.174)


3.6 Oscillating reactions* 67<br />

I<br />

Figure 3.5: Autocatalysis.<br />

• Ascanbeseenfromthesigmoid(-shaped) concentration-time profile of [B] in<br />

Fig. 3.5, the rate of increase of [B] increases up to the inflection point ∗ .<br />

• This curve is typical for a population increase from a low plateau to a new (higher)<br />

plateau after a favorable change of the environment due to, e.g., the availability<br />

of more food, a technical revolution, a reduction of mortality by a medical<br />

breakthrough, ...<br />

Many interesting games which can be played based on these ideas are described in (Eigen<br />

1975).<br />

3.6.2 <strong>Chemical</strong> oscillations<br />

The presence of one or more autocatalytic steps in a reaction mechanism can lead to<br />

chemical oscillations.<br />

a) The Lotka mechanism<br />

The origin of chemical oscillations can be understood by considering the reaction mechanism<br />

proposed by Lotka in 1910 (see Lotka 1920):<br />

(3.175)<br />

A+X 1<br />

→ X+X (1)<br />

X+Y 2<br />

→ Y+Y (2)<br />

Y<br />

<br />

→<br />

3<br />

Z (3)


3.6 Oscillating reactions* 68<br />

• Both reactions (1) and (2) are autocatalytic.<br />

• The Lotka mechanism illustrates the principle, but it does not correspond to an<br />

existing chemical reaction system.<br />

• A real oscillating reaction is the Belousov-Zhabotinsky reaction; this reaction may<br />

be described using a more complicated mechanism (see below).<br />

In order to model the resulting chemical oscillations, we assume that the reaction takes<br />

place in a flow reactor, which is constantly supplied with new A such that the concentration<br />

of A stays constant ([A] = [A] 0<br />

), while the product Z is constantly removed.<br />

I<br />

Rate equations and steady-state solutions:<br />

[X]<br />

<br />

[Y]<br />

<br />

=+ 1 [A] 0<br />

[X] − 2 [X] [Y] (3.176)<br />

=+ 2 [X] [Y] − 3 [Y] (3.177)<br />

Steady-state solutions:<br />

[X]<br />

<br />

[Y]<br />

<br />

=+ 1 [A] 0<br />

[X] − 2 [X] [Y] = 0 (3.178)<br />

=+ 2 [X] [Y] − 3 [Y] = 0 (3.179)<br />

Dividing these equations by [X] and [Y], respectively, we find<br />

2 [Y] <br />

= 1 [A] 0<br />

(3.180)<br />

2 [X] <br />

= 3 (3.181)<br />

Since [A] = [A] 0<br />

, the steady state solutions for [X] and [Y] are independent of time!<br />

I Displacements from steady-state solutions: What happens if the concentrations<br />

are displaced from the steady-state values by small amounts and ?<br />

(1) Ansatz:<br />

[X] = [X] <br />

+ (3.182)<br />

[Y] = [Y] <br />

+ (3.183)


3.6 Oscillating reactions* 69<br />

(2) Rate equations:<br />

<br />

=+ 1 [A] 0<br />

([X] <br />

+ ) − 2 ([X] <br />

+ )([Y] <br />

+ ) (3.184)<br />

=+ 1 [A] 0<br />

[X] <br />

+ 1 [A] 0<br />

− 2 [X] <br />

[Y] | {z }<br />

= 1 [A] 0<br />

[X] <br />

− 2 [X] <br />

− 2 [Y] <br />

− 2 <br />

| {z }<br />

= 1 [A] 0<br />

<br />

= − 2 [X] <br />

− 2 (3.185)<br />

<br />

=+ 2 ([X] <br />

+ )([Y] <br />

+ ) − 3 ([Y] <br />

+ ) (3.186)<br />

= 2 [X] <br />

[Y] <br />

+ 2 [X] <br />

+ 2 [Y] <br />

<br />

+ 2 − |{z} 3 [Y] <br />

− 3 |{z}<br />

= 2 [X] <br />

= 2 [X] <br />

=+ 2 [Y] <br />

+ 2 (3.187)<br />

(3) Simplified DE’s for small displacements: We may neglect the terms containing<br />

. Thus, we obtain two coupled first-order DE’s<br />

<br />

= − 2 [X] <br />

(3.188)<br />

<br />

=+ 2 [Y] <br />

(3.189)<br />

which can be converted to a single second-order DE by the ansatz<br />

·<br />

= − (3.190)<br />

·<br />

=+ (3.191)<br />

Second-order DE:<br />

··<br />

= − · = − (3.192)<br />

Formal solution of this second-order DE:<br />

() =[X()] − [X] <br />

= 1 + 2 − (3.193)<br />

(4) Eigenvalues: To determine the eigenvalue belonging to the linear DE’s, we have<br />

to solve the determinant<br />

¯<br />

¯0 − −<br />

+ 0 − ¯ − 2 [X] <br />

¯ 2 [Y] <br />

¯ =0 (3.194)<br />

y<br />

2 + 2 [X] <br />

[Y] <br />

=0 (3.195)<br />

y The solutions for are purely imaginary:<br />

12 = ± ¡ ¢<br />

2 2 12<br />

[X] <br />

[Y] <br />

(3.196)<br />

= ± ( 1 3 [A] 0<br />

) 12 (3.197)


3.6 Oscillating reactions* 70<br />

(5) Solution for ():<br />

() = 1 + 2 − = 0 cos (3.198)<br />

with<br />

=( 1 3 [A] 0<br />

) 12 (3.199)<br />

This means that if the system is displaced from its steady-state, the species concentrations<br />

will start to oscillate and the displacements may even grow in magnitude<br />

until the amplitude reaches the limiting value 0 (see limit cycle diagram in<br />

Fig. 3.6).<br />

I<br />

Figure 3.6: Limit cycle diagram for chemical oscillations according to the Lotka mechanism.<br />

b) The Brusselator 32<br />

Reaction scheme:<br />

<br />

A →<br />

1<br />

X (1)<br />

B+X 2<br />

→ R+Y (2)<br />

Y+2X (3.200)<br />

3<br />

→ 3X (3)<br />

<br />

X →<br />

4<br />

S (4)<br />

32 The name ‘Brusselator’ stems from the workplace of its inventor Prigogine (Brussels). The model<br />

does not correspond to a real chemical system.


3.6 Oscillating reactions* 71<br />

Here, A and B are reactants, R and S are final products, and X and Y are intermediates.<br />

The autocatalytic reaction is reaction (3).<br />

Rate equations (using “dimensionless time” = 4 ) and steady-state concentrations:<br />

[X]<br />

<br />

[X]<br />

<br />

Steady-state concentrations:<br />

= ·<br />

[X] = 1 [A] − 2 [B] [X] + 3 [X] 2 [Y] − 4 [X]<br />

4<br />

(3.201)<br />

= ·<br />

[Y] = 2 [B] [X] − 3 [X] 2 [Y]<br />

4<br />

(3.202)<br />

[X] <br />

[Y] <br />

= 1 [A]<br />

4<br />

(3.203)<br />

= 2 4 [B]<br />

1 3 [A]<br />

(3.204)<br />

We now investigate the effect of small displacements from steady-state by [X] and<br />

[Y]. The reaction rates respond to these displacements according to<br />

⎛ ⎞<br />

⎛ ⎞<br />

⎝ [X]<br />

·<br />

⎠ [X] and ⎝ [Y]<br />

·<br />

⎠ [Y] (3.205)<br />

[X]<br />

[Y]<br />

y<br />

• The steady-state is stable if<br />

⎛ ⎞<br />

⎝ [X]<br />

·<br />

⎠<br />

[X]<br />

• The steady-state is unstable if<br />

⎛ ⎞<br />

⎝ [X]<br />

·<br />

⎠<br />

[X]<br />

• Stability criterion:<br />

⎛ ⎞<br />

⎝ [X]<br />

·<br />

⎠<br />

[X]<br />

<br />

<br />

⎛ ⎞<br />

+ ⎝ [Y]<br />

·<br />

⎠<br />

[Y]<br />

<br />

<br />

<br />

⎛ ⎞<br />

+ ⎝ [Y]<br />

·<br />

⎠<br />

[Y]<br />

⎛ ⎞<br />

+ ⎝ [Y]<br />

·<br />

⎠<br />

[Y]<br />

<br />

<br />

<br />

0 (3.206)<br />

≥ 0 (3.207)<br />

= 2 []<br />

− 2 1 3 [] 2<br />

− 1 (3.208)<br />

4 4<br />

2<br />

c) Belousov-Zhabotinsky reaction<br />

Oxidation of malonic acid to CO 2 by bromate, catalyzed by cerium ions in acidic solution:<br />

2BrO − 3 +3CH 2 (COOH) 2<br />

+2H + −→ 2BrCH(COOH) 2<br />

+3CO 2 +4H 2 O (3.209)<br />

The reaction is probed by addition of the redox indicator ferroin/ferriin: The color varies<br />

between red and blue depending on the concentrations of Ce 4+ and Ce 3+


3.6 Oscillating reactions* 72<br />

I<br />

Figure 3.7: Mechanism of the Belousov-Zhabotinsky reaction.<br />

I<br />

Figure 3.8: Phases I - III of the Belousov-Zhabotinsky reaction.


3.6 Oscillating reactions* 73<br />

I<br />

Figure 3.9: Key steps of the Belousov-Zhabotinsky reaction.<br />

d) The Oregonator 33<br />

TheOregonatormodelcanbeusedtodescribetheBelousov-Zhabotinsky reaction.<br />

I<br />

Schematic model and kinetic equations:<br />

A+Y 1<br />

→ X+R (1)<br />

X+Y 2<br />

→ 2R (2)<br />

A+X 3<br />

→ 2X+2Z (3)<br />

(3.210)<br />

B+Z 5<br />

→ 1 2 Y (5)<br />

2X<br />

<br />

→<br />

4<br />

A +R (4)<br />

with A=BrO − 3 B=CH 2 (COOH) 2 + BrCH(COOH) 2 R=HOBr X=HBrO 2 <br />

Y=Br − Z=Ce 4+ .<br />

Using dimensionless time ( = 5 [B] ), we obtain<br />

33 The name ‘Oregonator’ stems from the workplace of its inventor Noyes (University of Oregon).


3.6 Oscillating reactions* 74<br />

[X]<br />

<br />

[Y]<br />

<br />

[Z]<br />

<br />

= 1 [A] [Y] − 2 [X] [Y] + 3 [A] [X] − 2 4 [X 2 ]<br />

5 [B]<br />

− 1 [A] [Y] − 2 [X] [Y] + 1<br />

=<br />

2 5 [B] [Z]<br />

5 [B]<br />

= 2 3 [A] [X] − 5 [B] [Z]<br />

5 [B]<br />

(3.211)<br />

(3.212)<br />

(3.213)<br />

I Figure 3.10: Numerical integration of the Oregonator model (rate constants in<br />

lmol −1 s −1 : 1 =0005, 2 =10, 3 =10, 4 =10, 5 = 0004; initial concentrations<br />

in mol l −1 : [A] = 001, [B] = 10, [X] 0<br />

=00, [Y] 0<br />

=00, [Z] 0<br />

=000025.)


3.6 Oscillating reactions* 75<br />

3.7 References<br />

Bader 1983 G. Bader, P. Deuflhard, Numerische Mathematik 41, 373 (1985)<br />

Deuflhard 1983 P. Deuflhard, Numerische Mathematik 41, 399 (1985).<br />

Deuflhard 1985 P. Deuflhard, SIAM Review 27, 505 (1985).<br />

Eigen 1975 M. Eigen, R. Winkler, Das Spiel, Piper, München, 1975.<br />

Gear 1971a C.W.Gear,Commun.Am.Comput.Mach.14, 1976 (1971).<br />

Gear 1971b C. W. Gear, Numerical Initial Value Problems in Ordinary Differential<br />

Equations, Prentice-Hall, Englewood Cliffs, 1971.<br />

Houston 1996 P. L. Houston, <strong>Chemical</strong> <strong>Kinetics</strong> and Reaction Dynamics, McGraw-<br />

Hill, Boston, 2001.<br />

Jost 1974 W. Jost, in <strong>Physical</strong> <strong>Chemistry</strong>: An Advanced Treatise, Ed.:W.Jost,Vol.<br />

VIa, Academic Press, New York, 1974.<br />

Lotka 1920 A. J. Lotka, Undamped Oscillations Derived from the Law of Mass Action,<br />

J. Am. Chem. Soc. 44, 1595 (1920).<br />

Press 1992 W. H. Press, S. A. Teukolsky, W. T. Vetterling, B. P. Flannery, Numerical<br />

Recipes in Fortran, Cambridge University Press, Cambridge, 1992.<br />

Scott 1994 S. K. Scott, Oscillations, Waves, and Chaos in <strong>Chemical</strong> <strong>Kinetics</strong>, Oxford<br />

Science Publications, Oxford, 1994.<br />

Stoer 1980 J. Stoer, R. Bulirsch, Introduction to Numerical Analysis, Springer,New<br />

York, 1980.


3.7 References 76<br />

4. Experimental methods<br />

In this section we consider experimental techniques for determining rate coefficients.<br />

We’ll get to faster reactions as we go down the list.<br />

Special techniques used in quantum state-resolved reaction dynamics, photodissociation,<br />

or transition state spectroscopy (molecular beams, state specific detection, photofragment<br />

translational spectroscopy, photofragment imaging, other pump-probe techniques,<br />

fluorescence up-conversion, fs-CARS, fs-DFWM) will be considered later (in advanced<br />

courses), but we’ll briefly touch photochemistry and “femtochemistry”.<br />

4.1 End product analysis<br />

• firststepforsettingupreactionmechanisms,<br />

• all convenient analytical techniques (HPLC, GC, GC-MS, MS, UV, FTIR, . . . ),<br />

• Problem: No “direct” kinetic information because of lack of time resolution.<br />

I<br />

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


4.2 Modulation techniques 77<br />

4.2 Modulation techniques<br />

• especially for photo-induced reactions.<br />

• photolysis light has to be modulated at a frequency that is comparable to that of<br />

the reaction y timescaleof10’sofmsandlonger.<br />

• kinetic information is derived from the “in-phase” and the “in-quadrature” components<br />

of the reactant and product concentrations as function of modulation<br />

frequency.<br />

I Figure 4.2:


4.3 The discharge flow (DF) technique 78<br />

4.3 The discharge flow (DF) technique<br />

DF technique is used for gas phase reactions on ms timescale:<br />

• laminar flow along tubular reactor:<br />

∆ = ∆<br />

<br />

with =meanflow speed, ∆ = distance along the flow tube.<br />

(4.1)<br />

• y first-order reaction gives exponential decay of concentration along ∆.<br />

• in reality ⇒ Hagen-Poisseuille flow (parabolic flow profile),<br />

• but fast radial diffusion of reactants (at pressures of 1 to 10 mbar) gives “plug<br />

shaped” radial concentration profiles despite parabolic flow profile.<br />

• more complicated, when “back diffusion” and radial diffusionhavetobetaken<br />

into account (at higher pressures).<br />

• problem: heterogeneous reactions lead to first-order decay of reactive radicals<br />

even in the absence of a reactant.<br />

Combinations with numerous highly sensitive detection techniques, for instance:<br />

• mass spectrometry (MS): universal, but often low sensitivity; problems with fragmentationofmoleculesintheionsource,<br />

• laser induced fluorescence (LIF): excitation to electronically excited state followed<br />

by relaxation by spontaneous emission. very sensitive (e.g., for OH ( 2 Σ ← 2 Π)<br />

detection limit below 10 6 molecules cm 3 ),<br />

• electron spin resonance (ESR): for paramagnetic atoms by exploiting the Zeeman<br />

effect depending on magnetic field <br />

= 0 + (4.2)<br />

∆ =1y ∆ = (4.3)<br />

ESR transitions are in the microwave region (X band: 9GHz).<br />

• laser magnetic resonance (LMR): very sensitive; based on observation of rotational<br />

transitions between Zeeman-shifted rotational levels of paramagnetic radicals<br />

( rot = rotational constant).<br />

00 = rot ( +1)+ 00<br />

00 (4.4)<br />

0 = rot ( +1)+ 0 0 (4.5)<br />

∆ = ±1 ∆ =0 ±1 : (4.6)<br />

y 0 = rot ( 00 +1)( 00 +2)+ 0 0 (4.7)<br />

y ∆ = =2 rot ( 00 +1)+ ( 0 0 − 00<br />

00 ) (4.8)<br />

LMR transitions are in the FIR region (100 m ≤ ≤ 2000 m).


4.3 The discharge flow (DF) technique 79<br />

I<br />

Figure 4.3: Schematic diagram of a DF/MS combination.<br />

I<br />

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


4.3 The discharge flow (DF) technique 80<br />

I<br />

Figure 4.5: Principles of Electron Spin Resonance (ESR) and Laser Magnetic Resonance<br />

(LMR).<br />

I<br />

Figure 4.6: Laser Magnetic Resonance (LMR).


4.3 The discharge flow (DF) technique 81<br />

I<br />

Figure 4.7: Schematic diagram of a DF/LMR combination<br />

I<br />

Figure 4.8: Flow reactor with laser flash photolysis and time-resolved mass spectrometric<br />

detection of transient reactants and products.


4.4 Flash photolysis (FP) 82<br />

4.4 Flash photolysis (FP)<br />

• “pump-probe” technique:<br />

<strong>—</strong> generation of transient species by flash photolysis using fast flashlampor<br />

pulsed lasers (“pump”),<br />

<strong>—</strong> fast detection of transient species by UV absorption spectrometry, laser induced<br />

fluorescence (LIF), cavity ringdown spectroscopy (CRDS) (”probe”).<br />

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

I<br />

Figure 4.9: Typical “Pump-Probe” set-up for kinetics studies by pulsed laser photolysis<br />

and detection by laser-induced fluorescence.<br />

I<br />

Cavity Ringdown Spectroscopy (CRDS):<br />

• CRDS is an ideal method for absorption spectroscopy with pulsed lasers despite<br />

of their large (±5 %) amplitude fluctuations.<br />

• The absorption measurement is transformed into the time domain:<br />

<strong>—</strong> time dependence of signal transmitted through a resonantor:<br />

= 0 exp (−) (4.9)


4.4 Flash photolysis (FP) 83<br />

<strong>—</strong> decay of intensity for pass through cavity (length , mirrorreflectivity ,<br />

absorption path length ):<br />

µ <br />

<br />

1 − <br />

= −<br />

<br />

−<br />

<br />

(4.10)<br />

| {z } | {z }<br />

due to reflection losses at mirrors due to absorption<br />

<strong>—</strong> with<br />

y<br />

<strong>—</strong> empty cavity ( =0):<br />

<strong>—</strong> with absorber ( 6= 0):<br />

= (4.11)<br />

<br />

= − [(1 − )+] (4.12)<br />

<br />

−1 (1 − )<br />

0 =<br />

<br />

(4.13)<br />

−1 =<br />

[(1 − )+]<br />

<br />

(4.14)<br />

<strong>—</strong> The difference of the ringdown times measures the absorption coefficient<br />

and is related to absorber concentration [A] by<br />

• Performance of a CRD spectrometer:<br />

<strong>—</strong> mirrors:<br />

<strong>—</strong> measuredemptycavitydecaytime:<br />

<strong>—</strong> effective absorption pathlength (1 m cell):<br />

• applications to chemical kinetics:<br />

−1 − −1<br />

0 = = [A] (4.15)<br />

=99999% (4.16)<br />

0 ≈ 300 s (4.17)<br />

eff =90km (4.18)<br />

<strong>—</strong> slow reactions: pump-probe scheme with variable time delay.<br />

first-order reaction with rate constant 1 :<br />

[A] = [A] 0<br />

− 1<br />

(4.19)<br />

y<br />

ln ¡ ¢<br />

−1 − −1<br />

0 = − 1 (4.20)<br />

<strong>—</strong> fast reactions: kinetics and ringdown occur on the same time scale. y<br />

(extended) <strong>Kinetics</strong> and Ringdown model ((e)SKaR); see Brown2000 and<br />

Guo2003 for details.


4.4 Flash photolysis (FP) 84<br />

I Figure 4.10:<br />

I Figure 4.11:


4.5 Shock tube studies of high temperature kinetics 85<br />

4.5 Shock tube studies of high temperature kinetics<br />

• Detection of atoms by atomic resonance absorption spectrometry (ARAS),<br />

• Detection of small radicals by frequency modulation absorption spectrometriy<br />

(FM-AS)<br />

I Figure 4.12:<br />

I Figure 4.13:


4.6 Stopped flow studies 86<br />

4.6 Stopped flow studies<br />

• the analogue of the flow tube technique for liquid phase reactions.<br />

• applications mainly kinetics of enzyme reactions, studies of protein folding dynamics<br />

I Figure 4.14:


4.7 NMR spectroscopy 87<br />

4.7 NMR spectroscopy<br />

• line coalescence<br />

• good for interconversion of two conformers (“isomerization”)<br />

I Figure 4.15:


4.8 Relaxation methods 88<br />

4.8 Relaxation methods<br />

• application to fast reactions in the liquid phase (Eigen)<br />

• -jump, -jump, -jump, . . . :<br />

(1) ⇒ displacement from equilibrium<br />

(2) ⇒ measurements of relaxation into new equilibrium state by transient absorption<br />

spectroscopy, conductivity, . . .


4.9 Femtosecond spectroscopy 89<br />

4.9 Femtosecond spectroscopy<br />

I<br />

Figure 4.16: Femtosecond pump-probe spectroscopy for the investigation of ultrafast<br />

reactions (time resolution down to the order of 10 fs).<br />

I<br />

Figure 4.17: Signals in broadband femtosecond transient absorption spectroscopy.


4.9 Femtosecond spectroscopy 90<br />

I Figure 4.18: Broadband femtosecond transient absorption spectroscopy of photochromic<br />

molecular switches in SFB 677.<br />

I<br />

Figure 4.19: Broadband femtosecond transient absorption spectroscopy of Disperse<br />

Red 1.


4.9 Femtosecond spectroscopy 91<br />

I Figure 4.20: Broadband femtosecond transient absorption spectroscopy of Furyl<br />

Fulgides.<br />

I<br />

Figure 4.21: Ultrafast Electronic Deactivation in DNA Building Blocks.


4.9 Femtosecond spectroscopy 92<br />

I<br />

Figure 4.22: Femtosecond fluorescence up-conversion data for DNA strands.<br />

I<br />

Figure 4.23: Ultrafast Electronic Deactivation in DNA Building Blocks: d(pApA) as<br />

Model System.


4.9 Femtosecond spectroscopy 93<br />

4.10 References<br />

Busch 1999 K. W. Busch, M. A. Busch, Cavity-Ringdown Spectroscopy, ACSSymposium<br />

Series, Vol 720, American <strong>Chemical</strong> Society, Washington, 1999.<br />

Brown 2000 S.S.Brown,A.R.Ravishankara,H.Stark,J.Phys.Chem.A104, 7044<br />

(2000).<br />

Friebolin 1999 H.Friebolin, Ein- und zweidimensionale NMR-Spektroskopie, Wiley-<br />

VCH, Weinheim, 1999.<br />

Guo 2003 Y. Guo, M. Fikri, G. Friedrichs, F. Temps, An Extended Simultaneous <strong>Kinetics</strong><br />

and Ringdown Model: Determination of the Rate Constant for the Reaction<br />

SiH 2 +O 2 .Phys.Chem.Chem.Phys.5, 4622 (2003).<br />

OKeefe 1988 A.O’Keefe,D.A.G.Deacon,Rev.Sci.Instrum.59, 2544 (1988).


4.10 References 94<br />

5. Collision theory<br />

So far, we have considered more or less complex chemical reactions following more or<br />

less complicated rate laws, but we have only considered the rate coefficients appearing<br />

in the rate laws as empirical quantities, which had to be determined experimentally.<br />

We shall now ask the question, whether we can rationalize and predict those rate coefficients<br />

using theoretical models. Towards these ends, we have to look at the reactions<br />

at a microscopic, molecular level. In particular, we have to<br />

• consider the relation between the microscopic (individual molecule) properties and<br />

the macroscopic (thermally averaged) rate quantities (rate constants, product<br />

branching ratios)<br />

• develop models for the microscopic (molecular level) progress of individual elementary<br />

reactions, including the different molecular degrees of freedom (quantum<br />

states).<br />

• develop sufficiently simplified models averaging over the molecular degrees of<br />

freedom which still allow us to make practically useful, relevant, and sufficiently<br />

accurate predictions.<br />

We shall do so in the next four chapters.<br />

I Collision theory: To begin with, this chapter considers bimolecular reactions in the<br />

gas phase. These obviously require a collision between two molecules as condition for a<br />

reaction to take place.<br />

• We start with the hard sphere collision model of structureless particles as the<br />

simplest model for bimolecular reactions. Although this model has two major<br />

flaws, it leads to the right result because of a compensation of errors.<br />

• We then consider results of kinetic gas theory, which allow us to properly take<br />

into account the molecular speed distributions.<br />

• We apply the hard sphere collision model to understand transport processes in the<br />

gas phase (diffusion, heat conductivity, viscosity).<br />

• We then proceed with a correct derivation of advanced collision theory models<br />

starting from differential cross sections (which can be measured in crossed molecular<br />

beam experiments). Advanced collision theory is the method of choice for fast<br />

gas phase reactions which can be studied in detail by molecular beam experiments<br />

and handled by quantum theory.


5.1 Hardspherecollisiontheory 95<br />

5.1 Hard sphere collision theory<br />

5.1.1 The gas kinetic collision frequency<br />

I<br />

The gas kinetic collision frequency as upper limit for rate constants of bimolecular<br />

reactions: The gas kinetic collision frequency constitues an upper limit for the<br />

rate constant of bimolecular gas phase reactions. We assume that the molecules A and<br />

B are structureless hard spheres (radii , ) moving around with fixedmeanspeeds<br />

and and look at the number of collisions between A and B per unit time.<br />

I<br />

Figure 5.1: Derivation of the hard sphere gas kinetic collision frequency.<br />

I<br />

Discussion of the molecular quantities affecting the rate constant:<br />

• collision parameter (= projected distance between centers of gravity):<br />

(5.1)<br />

• maximal collision parameter = max value leading to a collision max :<br />

≤ max = + = (5.2)<br />

• collision cross section = 2 (= reactive cross section) 34 :<br />

= 2 = ( + ) 2 (5.3)<br />

34 Deutsch: reaktiver Wirkungsquerschnitt.


5.1 Hardspherecollisiontheory 96<br />

• mean speed of A and B:<br />

=<br />

=<br />

r r<br />

8 8<br />

=<br />

(5.4)<br />

<br />

r r <br />

8 8<br />

=<br />

(5.5)<br />

<br />

• mean relative speed (2 different molecules, i.e., A 6= B): 35<br />

=<br />

r 8 <br />

<br />

(5.9)<br />

with the reduced mass<br />

=<br />

<br />

+ <br />

(5.10)<br />

• mean relative speed (single type of molecules, i.e., A = B):<br />

y<br />

=<br />

<br />

+ <br />

= 1 2 (5.11)<br />

= √ 2 ×<br />

r<br />

8 <br />

<br />

(5.12)<br />

I<br />

Hard sphere gas kinetic collision frequency:<br />

• We want to determine the number of collisions of a single molecule A with<br />

molecules B in time . This is given by the product of the (volume that A has<br />

swept in ) × (number density of B):<br />

<strong>—</strong> volume of cylinder that A has swept in time :<br />

volume = cross section × (5.13)<br />

y<br />

2 × (5.14)<br />

35 The mean relative speed of two molecules with velocities u and u is found by looking at the<br />

square of the difference speed<br />

2 = |u − u | 2 = | | 2 + | | 2 − 2 | || | cos (5.6)<br />

The third term averaged over cos is 0, therefore<br />

Since and<br />

2 = | | 2 + | | 2 = 3 <br />

<br />

+ 3 <br />

= 3 ( + )<br />

= 3 <br />

<br />

³<br />

2´12 differ only by a constant factor, we also find that<br />

=<br />

s<br />

8 <br />

<br />

(5.7)<br />

(5.8)


5.1 Hardspherecollisiontheory 97<br />

<strong>—</strong> number density of B<br />

(5.15)<br />

<strong>—</strong> Number of collisions of a single A in time : ⇒ number of molecules B in<br />

that volume<br />

= 2 × × (5.16)<br />

• Number of all collisions between A and B:<br />

= 2 × × (5.17)<br />

• Modification for collisions between identical molecules (collisions must not be<br />

counted twice):<br />

= 1 2 × 2 × × (5.18)<br />

=<br />

√<br />

2<br />

2 × 2 × × (5.19)<br />

• Comparison with rate expression:<br />

y<br />

− <br />

<br />

= − <br />

<br />

= × = (5.20)<br />

• Expression for rate constant for collisions A + B:<br />

<br />

max = 2 × = × (5.21)<br />

y<br />

µ 12<br />

max = 2 8 <br />

×<br />

(5.22)<br />

<br />

• Modification for collisions between identical molecules (A + A):<br />

max = 1 2 2 × = √ 1 µ 12<br />

2 8 <br />

×<br />

(5.23)<br />

2 <br />

Theseexpressionsgivetheupper limits for the rate constants of bimolecular<br />

reactions. The upper limits are reached only if<br />

• there are no energetic restrictions (no energy barriers),<br />

• all collisions lead to reaction (no steric constraints),<br />

• molecules can be approximated as hard spheres (unrealistic, usually one has <br />

),<br />

• all molecules collide with the same relative speed (in reality there is a distribution<br />

of speeds).


5.1 Hardspherecollisiontheory 98<br />

I<br />

Mean free pathlength:<br />

• The mean free pathlength is the distance that a molecule flies between two successive<br />

collisions:<br />

y<br />

=<br />

distance<br />

# of collisions = distance/time<br />

# of collisions/time = <br />

(5.24)<br />

<br />

=<br />

1<br />

2 × <br />

∝ 1 <br />

• Modification for collisions between identical molecules (A + A):<br />

(5.25)<br />

= 1 √<br />

2<br />

1<br />

2 × <br />

(5.26)<br />

I<br />

Time between two collisions:<br />

1<br />

<br />

=<br />

1<br />

2 × <br />

(5.27)<br />

I<br />

Example 5.1: Collision frequency for O 2 -N 2 @ =298K= 1000 mbar:<br />

Molecular parameters:<br />

2 = 2 2 =3467 Å (5.28)<br />

2 = 2 2 =3798 Å (5.29)<br />

y =36 Å =36 × 10 −10 m (5.30)<br />

= 2 =41Å 2 (5.31)<br />

28 × 32<br />

=<br />

28 + 32 × 10−3 kg mol × 1<br />

<br />

s<br />

<br />

(5.32)<br />

=<br />

8<br />

<br />

=650m s (5.33)<br />

(but (O 2 )=444m s) (5.34)<br />

Rate constant:<br />

µ 12<br />

max = 2 8 <br />

×<br />

<br />

(5.35)<br />

=26 × 10 −10 cm 3 s (5.36)<br />

=16 × 10 14 cm 3 mol s (5.37)<br />

Mean free pathlength:<br />

=1000mbar (5.38)<br />

=40 × 10 −5 mol cm 3 (5.39)<br />

=25 × 10 25 m 3 (5.40)<br />

=36 × 10 −10 m (5.41)<br />

=1× 10 −7 m (5.42)


5.1 Hardspherecollisiontheory 99<br />

Rule-of-thumb:<br />

Time between collisions:<br />

=03mbary =03mm (5.43)<br />

=1000mbar (5.44)<br />

1 1<br />

=<br />

2 × <br />

(5.45)<br />

=15 × 10 −10 s (5.46)<br />

≈ 01ns (5.47)<br />

¤<br />

5.1.2 Allowance for a threshold energy for reaction<br />

I The threshold energy E 0 : The majority of reactions exhibit a threshold energy, below<br />

which the reaction does not occur (quantum mechanical tunneling is excluded here).<br />

The molecules must have a minimum energy min to overcome this reaction threshold energy<br />

0 . For our structureless spherical molecules, we consider only translational energy<br />

: 36 0 : no reaction (5.48)<br />

0 : reaction (5.49)<br />

I Boltzmann distribution: Fraction of molecules with translational energy 0 :<br />

µ<br />

( 0 )<br />

=exp − µ<br />

0<br />

=exp − <br />

0<br />

(5.50)<br />

<br />

<br />

<br />

I<br />

Resulting rate constant:<br />

or<br />

µ 12<br />

max = 2 8 <br />

×<br />

× − 0 <br />

(5.51)<br />

<br />

µ 12<br />

max = 2 8 <br />

×<br />

× − 0<br />

<br />

(5.52)<br />

I Arrhenius activation energy: Applying the definition of the apparent (Arrhenius)<br />

activation energy ,wefind that<br />

= 2 ln <br />

<br />

= 0 + 1 (5.53)<br />

2<br />

As we see, is not constant but (slightly) -dependent: The slope of a plot of ln <br />

max<br />

vs. −1 increases as →∞(or −1 → 0). The Arrhenius preexponential factor <br />

would not be constant as well, it is also slightly -dependent.<br />

36 We use and 0 to indicate the energies in molecular units, and and 0 for the respective<br />

values in molar units.


5.1 Hard sphere collision theory 100<br />

5.1.3 Allowance for steric hindrance<br />

I The sterical factor p: In addition, even if 0 , not every collision will lead to<br />

a reaction, because real molecules are not structureless: A reaction requires a specific<br />

orientation of the molecules. This is taken into account by introducing a steric factor<br />

:<br />

µ 12<br />

max = × 2 8 <br />

×<br />

× − 0<br />

(5.54)


5.2 Kinetic gas theory 101<br />

5.2 Kinetic gas theory<br />

5.2.1 The Maxwell-Boltzmann speed distributions of gas molecules in 1D and 3D<br />

I<br />

Figure 5.2: Experimental measurement of molecular speed distribution using a rotating<br />

disk velocity selector and an effusive molecular beam (orrifice diameter ¿ ).<br />

I<br />

Rotating disk velocity selector:<br />

• Flight time of molecules with velocity between two rotating disks (distance ):<br />

1 = <br />

(5.55)<br />

• Time delay of second slit with respect to first slit ( =angle, = angular velocity):<br />

2 = <br />

(5.56)<br />

• Transmission only if<br />

y<br />

1 = 2 (5.57)<br />

= × <br />

(5.58)<br />

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

speed distribution ( ) = ( )<br />

<br />

using this method gives


5.2 Kinetic gas theory 102<br />

I The 1D Maxwell-Boltzmann speed distribution (Fig. 5.3):<br />

r<br />

µ<br />

<br />

( ) =<br />

2 × exp − 2 <br />

2 <br />

<br />

(5.59)<br />

Note that this distribution is already normalized for<br />

Z ∞<br />

( ) =1 We check this<br />

by substituting<br />

r<br />

r<br />

r<br />

<br />

<br />

=<br />

2 y =<br />

2 2 <br />

y = (5.60)<br />

<br />

y<br />

Z ∞<br />

r Z ∞ r<br />

2 <br />

( ) =<br />

= 1 Z∞<br />

√ <br />

2 −2 −2 = 1 √<br />

√ =1 (5.61)<br />

<br />

−∞<br />

−∞<br />

−∞<br />

−∞<br />

I<br />

Figure 5.3: 1D Maxwell-Boltzmann speed distribution.


5.2 Kinetic gas theory 103<br />

I Velocity distribution in 3 dimensions: Since the velocities in direction are<br />

independent, the probability that a molecule has a velocity between and + ,<br />

and + ,and and + is given be the product of ( ) , ( ) ,<br />

and ( ) :<br />

( )<br />

<br />

= ( ) ( ) ( ) <br />

µ Ã<br />

32 <br />

=<br />

× exp − ¡ ¢ !<br />

2 + 2 + 2 <br />

(5.62)<br />

2 <br />

2 <br />

We are, however, not interested in the distribution of velocities pointing into a specific<br />

direction ( ), but in the probability that a molecule has a certain speed || in<br />

any direction, i.e., that the velocity vector ends somewhere on a sphere with radius<br />

q<br />

|| = 2 + 2 + 2 (5.63)<br />

To obtain this distribution, we must integrate over all polar angles and . This<br />

corresponds to replacing the volume element with the volume 4 2 of<br />

a spherical shell with radius and thickness y<br />

I The 3D Maxwell-Boltzmann speed distribution (Fig. 5.4):<br />

() =<br />

µ 32 <br />

<br />

× 4 2 × exp<br />

µ− 2 (5.64)<br />

2 <br />

2 <br />

Note that this distribution is also normalized for<br />

Z ∞<br />

() =1 We check this by<br />

substituting<br />

y<br />

Z ∞<br />

0<br />

r<br />

r<br />

µ 12 <br />

<br />

=<br />

2 y = 2 y = 2 <br />

<br />

<br />

(5.65)<br />

= 2 <br />

2 (5.66)<br />

2<br />

() =<br />

µ 32 <br />

× 4<br />

2 <br />

0<br />

Z ∞<br />

0<br />

0<br />

µ<br />

2 2 <br />

2 −2 <br />

12<br />

(5.67)<br />

µ 32 Z∞<br />

µ 32 1 1<br />

=4 × 2 −2 =4 × × 1 √ =1 (5.68)<br />

4


5.2 Kinetic gas theory 104<br />

I<br />

Figure 5.4: 3D Maxwell-Boltzmann speed distribution.<br />

I<br />

Notice the differences between the following quantities:<br />

• Value at maximum of ():<br />

• Mean speed:<br />

• Root mean square value:<br />

max =<br />

=<br />

r<br />

2 <br />

<br />

r<br />

8 <br />

<br />

³<br />

2´12 =<br />

r<br />

3 <br />

<br />

(5.69)<br />

(5.70)<br />

(5.71)<br />

I<br />

Table 5.1: Mean molecular speed values for some gases @ =298K<br />

gas [m s]<br />

H 2511<br />

H 2 1776<br />

He 1255<br />

O 2 444<br />

CO 2 379<br />

Hg 178<br />

238 U 163


5.2 Kinetic gas theory 105<br />

I Kinetic energy distribution function: We may substitute for in Eq. 5.64:<br />

= 1 2 2 (5.72)<br />

µ 12 2<br />

y =<br />

(5.73)<br />

<br />

y<br />

µ 12 2<br />

= × 1 µ 12 1<br />

2 × ( ) −12 = ( ) −12 (5.74)<br />

2<br />

µ 12 1<br />

y = ( ) −12 (5.75)<br />

2<br />

Insertion into Eq. 5.64 gives (see Fig. 5.5)<br />

( ) =<br />

y<br />

µ 32 µ <br />

µ 12 µ 12 <br />

2<br />

× 4 × − 1 1<br />

×<br />

(5.76)<br />

2 <br />

<br />

2 <br />

µ 32 1<br />

( ) =2<br />

( ) 12 − (5.77)<br />

<br />

I<br />

Figure 5.5: Maxwell-Boltzmann distribution function for the kinetic energy.<br />

The fraction of molecules with 0 is temperature dependent.


5.2 Kinetic gas theory 106<br />

5.2.2 The rate of wall collisions in 3D<br />

Earlier, we calculated the rate of wall collisions of the molecules of a gas with mass <br />

and number density in a container of volume and the resulting gas pressure using<br />

asimplified model by assuming that 1 r<br />

6 of all molecules have a fixed speed = 8 <br />

<br />

in + direction (cf. Fig. 5.6). In this simple picture, we neglected the 3D molecular<br />

velocity distribution. We obtained the correct result only due to a cancelation of two<br />

errors (assumption of fixed 1 and 3D distribution). In the following,<br />

6<br />

(1) we briefly review the simplified picture,<br />

(2) we then proceed to derive the correct results by appropriate averaging over the<br />

velocity/speed distribution.<br />

I<br />

Figure 5.6: Simplified model for the calculation of the pressure of an ideal gas.<br />

I<br />

Simplified model for the calculation of the pressure of an ideal gas:<br />

• In the time , all gas molecules (mass ) inthevolume + hit the wall<br />

area .<br />

• Number of wall collisions in time (with number density ):<br />

= × 1 × × (5.78)<br />

6


5.2 Kinetic gas theory 107<br />

• Momentum change due to wall collision, force on wall, and pressure:<br />

=2 × 1 (5.79)<br />

6<br />

• Gas pressure:<br />

y<br />

= = 1 <br />

= 1 × 2 × 1 (5.80)<br />

6<br />

= 1 3 2 (5.81)<br />

• Result for 1mol ( = <br />

<br />

with =6022 1 × 10 23 mol −1 and kin = 2<br />

2 ):<br />

= 1 3 2 = 2 3 kin = 2 3 × 3 2 = (5.82)<br />

I<br />

Correct derivation of the gas kinetic wall collision frequency:<br />

• Number of molecules with velocities in a volume of with =<br />

p <br />

2 + 2 + 2 (cf. Fig. 5.7) that will hit the area in the time :<br />

( )=| | ( )<br />

<br />

<br />

( )<br />

<br />

( )<br />

<br />

(5.83)<br />

• To find the number of all molecules hitting in , we have to integrate over all<br />

positive and over all (positive and negative) and , using the 1D-Maxwell-<br />

Boltzmann distributions ( ) = ( )<br />

<br />

,etc.:<br />

µ 32 <br />

= <br />

×<br />

2 <br />

Z ∞ Z ∞ Z ∞ µ<br />

| | exp − 2 <br />

2 <br />

0 −∞ −∞<br />

µ µ <br />

exp − 2 <br />

exp − 2 <br />

<br />

2 2 <br />

(5.84)<br />

We know that the distributions over and are normalized to 1. Hence, we<br />

only have to solve the integral over , which we do by the substitution<br />

µ 12 µ 12 µ 12 <br />

<br />

2 <br />

=<br />

y =<br />

y =<br />

(5.85)<br />

2 <br />

2 <br />

<br />

y<br />

Z ∞<br />

0<br />

µ <br />

| | exp − 2 <br />

=<br />

2 <br />

Z ∞<br />

0<br />

= 2 <br />

<br />

µ 2 <br />

<br />

Z ∞<br />

0<br />

12<br />

|| −2 µ 2 <br />

<br />

−2 = 2 <br />

<br />

1<br />

2 = <br />

<br />

12<br />

(5.86)<br />

(5.87)


5.2 Kinetic gas theory 108<br />

• Result: 37<br />

= <br />

µ 12 <br />

× <br />

2 <br />

µ 12 = × <br />

(5.88)<br />

2<br />

y<br />

wall = µ 12<br />

= × <br />

(5.89)<br />

2<br />

This can be rewritten using the mean molecular speed , because<br />

With<br />

we thus obtain the<br />

wall = × 1 4<br />

µ 16 <br />

2<br />

=<br />

12<br />

= × 1 4<br />

µ 8 <br />

<br />

• Final result for the gas kinetic collision frequency:<br />

µ 8 <br />

<br />

12<br />

(5.90)<br />

12<br />

(5.91)<br />

wall = 1 4 (5.92)<br />

I<br />

Figure 5.7: On the derivation of the gas kinetic wall collision frequency.<br />

µ 12 <br />

37 Notice that two of the three<br />

factors in Eq. 5.84 above are compensated by the<br />

2 <br />

µ 12 <br />

integrations over and ,onlyone<br />

factor remains.<br />

2


5.2 Kinetic gas theory 109<br />

I<br />

Correct derivation of the gas pressure and the ideal gas equation:<br />

• Likewise, we take the differential number of wall collisions for specific speeds(Eq.<br />

5.83) to compute the momentum change ( ) in time <br />

( )=2 | |×| | × ( ) ( ) ( ) <br />

(5.93)<br />

• Inserting again the 1D-Maxwell-Boltzmann distributions and integrating over all<br />

positive and all ,wefind the total momentum change in time :<br />

µ 32 <br />

= <br />

× 2×<br />

2 <br />

Z ∞ Z ∞ Z ∞ µ<br />

2 exp − 2 <br />

2 <br />

0 −∞ −∞<br />

µ µ <br />

exp − 2 <br />

exp − 2 <br />

<br />

2 2 <br />

• Pressure for = µ 12<br />

<br />

<br />

again using the substitution = <br />

2 <br />

y<br />

y<br />

= <br />

= <br />

<br />

= <br />

<br />

µ 12 <br />

× 2<br />

2 <br />

µ 1<br />

<br />

12<br />

2 2 <br />

<br />

Z ∞<br />

0<br />

Z ∞<br />

0<br />

(5.94)<br />

µ <br />

2 exp − 2 <br />

(5.95)<br />

2 <br />

2 −2 = <br />

(5.96)<br />

= (5.97)<br />

Thus, we have fixed our earlier simple derivations by appropriately integrating over the<br />

Maxwell-Boltzmann velocity distribution of the molecules.


5.3 Transport processes in gases 110<br />

5.3 Transport processes in gases<br />

5.3.1 The general transport equation<br />

I Definition of the flux −→ J of a quantity: We consider a generic transport property<br />

Γ (5.98)<br />

which differs in magnitude along the direction, i.e., has the gradient<br />

grad Γ (5.99)<br />

The flux of Γ throughanarea in time is defined as<br />

andcanbewrittenas<br />

−→ Γ = Γ<br />

<br />

(5.100)<br />

−→ Γ = − grad Γ (5.101)<br />

Kinetic gas theory allows us to derive a general equation for the proportionality constant<br />

.<br />

I<br />

Figure 5.8: Derivation of the general transport equation.


5.3 Transport processes in gases 111<br />

I The general transport equation: We consider the net flux of Γ through an area <br />

at = 0 that is transported by gas molecules coming from an area above at = 0 +<br />

and from below from an area above at = 0 − (where isthemeanfreepathlength).<br />

Denoting the respective transport quantity per molecule as Γ, these partial fluxes are<br />

given by the number of gas kinetic collisions hitting on ( = 0 ) times the Γ carried<br />

by each gas molecule (number density ):<br />

• Using the gas kinetic wall collision frequency wall = 1 (Eq. 5.92), we obtain<br />

4<br />

Γ + = 1 µ Γ<br />

µΓ<br />

4 × × 0 + (5.102)<br />

<br />

and<br />

• Net effect:<br />

Γ − = 1 4 × × µΓ 0 −<br />

µ Γ<br />

(5.103)<br />

<br />

Γ = Γ + − Γ − = 1 µ Γ<br />

2 × × (5.104)<br />

<br />

• Net flux (with − sign to account for the fact that the flux is directed along the<br />

downhill gradient of Γ):<br />

−→ Γ = Γ<br />

µ Γ<br />

= −1 2 <br />

with given by (see the derivation of wall above)<br />

=<br />

µ 8 <br />

<br />

(5.105)<br />

12<br />

(5.106)<br />

and, for only one type of gas molecules A (i.e., A - A collisons),<br />

= 1 √<br />

2<br />

1<br />

2 × <br />

(5.107)<br />

• If also varies along (diffusion), Eq. 5.105 may be recast by defining the overall<br />

transport quantity Γ 0 from the transport quantity per molecule Γ as<br />

Γ 0 = Γ (5.108)<br />

into the form<br />

−→ Γ = − 1 2 Ã<br />

!<br />

Γ 0<br />

<br />

(5.109)<br />

In the following, ³ we shall apply Eqs. 5.105 or 5.109 to determine the self-diffusion coefficient<br />

Γ =1 Γ 0 = <br />

´, the heat conductivity ¡ Γ = ¢ ,andtheviscosity ¡ ¢<br />

Γ = <br />

of gases.


5.3 Transport processes in gases 112<br />

5.3.2 Diffusion<br />

I<br />

Figure 5.9: The diffusion in a mixture A + B of one sort of molecules A against a<br />

time-independent concentration gradient <br />

<br />

is described by Fick’s first law.<br />

• Fick’s first law:<br />

−→ = µ <br />

<br />

= − <br />

<br />

(5.110)<br />

• In order to derive an expression for the self-diffusion coefficient (diffusion of an<br />

isotope of A in normal A), we make the ansatz<br />

Γ =1 Γ 0 = (5.111)<br />

y<br />

• Result:<br />

−→ = − 1 µ <br />

2 <br />

<br />

(5.112)<br />

= 1 (5.113)<br />

2<br />

• Fick’s second law (time-dependent concentration gradient):<br />

• Einstein’s diffusion law in 1D:<br />

• Einstein’s diffusion law in 3D:<br />

<br />

<br />

= 2 <br />

2 (5.114)<br />

∆ 2 =2 (5.115)<br />

∆ 2 =6 (5.116)


5.3 Transport processes in gases 113<br />

5.3.3 Heat conduction<br />

I<br />

Figure 5.10: Heat conduction is described by Fourier’s law.<br />

• Fourier’s law (for constant temperature gradient):<br />

−→ =<br />

µ <br />

= − <br />

(5.117)<br />

• Ansatz:<br />

y<br />

• Result:<br />

Γ = = internal energy per molecule (5.118)<br />

−→ <br />

= − 1 µ <br />

2 <br />

= − 1 µ µ <br />

<br />

2 <br />

= − 1 µ µ<br />

2 <br />

<br />

= − 1 µ <br />

2 <br />

<br />

<br />

(5.119)<br />

(5.120)<br />

(5.121)<br />

(5.122)<br />

= 1 2 (5.123)<br />

with = number density and = specific heat per molecule.


5.3 Transport processes in gases 114<br />

5.3.4 Viscosity<br />

I<br />

Figure 5.11: Newton’s law describes the inner friction in laminar flow.<br />

• Newton’s law:<br />

• Ansatz:<br />

µ <br />

<br />

= − <br />

<br />

(5.124)<br />

Γ = = momentum per molecule in direction (5.125)<br />

y<br />

• Result:<br />

−→ = − 1 µ <br />

2 ( )<br />

<br />

= 1 ( )<br />

= 1 <br />

= <br />

= 1 (5.126)<br />

2<br />

with = number density and = mass of molecule.


5.4 Advanced collision theory 115<br />

5.4 Advanced collision theory<br />

The hard sphere model is a very primitive model. In reality, we have to account for the<br />

following:<br />

• The reaction probability depends on exact details of the collision, i.e.,<br />

hard sphere model → ( ) (5.127)<br />

• Molecules are not hard spheres with = 2 but “soft”, i.e., the reactive cross<br />

section depends on the relative speed (or kinetic energy)<br />

→ () → ( ) (5.128)<br />

• The speed distribution follows the Maxwell-Boltzmann distribution, i.e.,<br />

→ () → ( ) (5.129)<br />

• Molecules in different quantum states behave differently, i.e.,<br />

→ ( ; ) (5.130)<br />

I Examples: The complexity of the problem is illustrated in Fig. 5.12 by two examples:<br />

D( )+H 2 ( 2 2 2 ) −→ HD( )+H( ) (5.131)<br />

OH( Ω )+H 2 ( 2 2 2 )<br />

→ H 2 O( 2 1 2 3 )+H( ) (5.132)<br />

As can be seen, the reaction rate is expected to depend on (speed, translational<br />

energy), (vibrational quantum numbers ( 1 = symmetric stretch, 2 = bend, 3 =<br />

antisymmetric stretch)), (rotational quantum numbers), Ω (fine structure<br />

state, electronic state). In addition, the products are scattered into the spatial angles<br />

.<br />

I<br />

Figure 5.12: Quantum state-specific elementary reactions.


5.4 Advanced collision theory 116<br />

We shall now derive the fundamental equation of collision theory that allows us to<br />

account for those effects.<br />

5.4.1 Fundamental equation for reactive scatteringincrossedmolecularbeams<br />

• Consider the following reactive scattering process (Fig. 5.13):<br />

A+B→ C (5.133)<br />

• IntensityofbeamofA:<br />

= × (5.134)<br />

Is this reasonable? ⇒ Look at the physical dimensions of × :<br />

[cm s] × £ molecules/ cm 3¤ = £ molecules cm 2 s ¤ =[intensity] (5.135)<br />

• Depletion of intensity of A due to scattering by B:<br />

= − () × × × <br />

↑<br />

↑<br />

↑<br />

<br />

<br />

(5.136)<br />

y rate equation:<br />

<br />

= −() (5.137)<br />

<br />

• Rate coefficient for a given value of :<br />

() =() (5.138)<br />

• Averaging over the speed distribution () leads to the fundamental equation<br />

of all improved collision theories:<br />

( )=h()i (5.139)<br />

i.e.,<br />

( )= ∞ R<br />

0<br />

() ( ) (5.140)<br />

• Frequently used approximation: Since the explicit form of () is usually not<br />

known, we write<br />

( )=hi h()i (5.141)<br />

with<br />

and<br />

hi =<br />

µ 8 <br />

<br />

12<br />

(5.142)<br />

h()i = -averaged (reactive) cross section (5.143)


5.4 Advanced collision theory 117<br />

I<br />

Figure 5.13: Derivation of the fundamental equation.<br />

5.4.2 Crossed molecular beam experiments<br />

I Differential and integral reactive cross sections: Experiments with crossed molecular<br />

beams allow us to measure differential and integral reactive cross sections<br />

(Herschbach 1987, Lee 1987).<br />

We consider the reaction<br />

( )+ ( ) −→ ( )+ ( ) (5.144)<br />

where denotes the speeds of the molecules and stands for the quantum numbers<br />

of their energy states (electronic, vibrational, rotational, nuclear spin). In an ideal<br />

experiment (which has yet to be performed), all of the reactants are defined by, e.g.,<br />

a specific excitation using laser, and the of the products are determined by some<br />

probe laser. The can be determined in a molecular beam experiment using a rotating<br />

disk velocity selector (see Fig. 5.14).


5.4 Advanced collision theory 118<br />

I<br />

Figure 5.14: Measurement of differential cross sections in crossed molecular beam<br />

experiments.<br />

I<br />

Differential and integral cross sections from crossed molecular beam experiments:<br />

• We define the differential cross section ( ) for scattering<br />

of products C into the solid angle element Ω at such that<br />

( ) Ω<br />

= ( )<br />

<br />

× × × ( ) × × ( ) × Ω (5.145)<br />

• Corresponding integral cross sections are obtain by integration over and .<br />

• Depending on the scattering process of interest, we distinguish between<br />

<strong>—</strong> elastic cross sections ( ): measure of the size of the molecules,<br />

<strong>—</strong> inelastic cross sections ( ):measureoftheefficiency of energy transfer<br />

processes,<br />

<strong>—</strong> reactive cross sections ( ): measure of reactivity.


5.4 Advanced collision theory 119<br />

I<br />

Figure 5.15: Measurement of differential cross sections in crossed molecular beam<br />

experiments.


5.4 Advanced collision theory 120<br />

5.4.3 From differential cross sections σ (...) to reaction rate constants k (T )<br />

I<br />

Figure 5.16: The strategy.<br />

I<br />

1 → 2:Integration over angles:<br />

0<br />

( )=<br />

Z 2 Z <br />

0<br />

0<br />

( ) 0<br />

( )sin (5.146)<br />

I 2 → 3:Summation over all product quantum states α 0 :<br />

( )= X 0<br />

0<br />

( ) (5.147)<br />

I 3 → 4:Thermal averaging over initial quantum states α:<br />

( )= X <br />

× ( ) (5.148)<br />

with the Boltzmann factor and the partition function <br />

=<br />

− <br />

P<br />

− <br />

= − <br />

<br />

(5.149)<br />

= X <br />

− <br />

(5.150)<br />

y<br />

( )= X <br />

( ) × − <br />

<br />

(5.151)


5.4 Advanced collision theory 121<br />

I 5 → 6:Thermal averaging over all ² :<br />

( )= X <br />

× ( )<br />

= X − <br />

× ( )<br />

<br />

<br />

Z<br />

− <br />

=<br />

× ( ) <br />

<br />

<br />

I<br />

3 → 5 and 4 → 6:Laplace Transformation:<br />

Z ∞<br />

( ) ∝ × ( ) × − (5.152)<br />

0<br />

This expression follows from the fundamental equation (Eq. 5.140) by substituting <br />

for . The exact transformation (giving the correct factors before the integral) will be<br />

carried out below (see Eq. 5.162).<br />

5.4.4 Laplace transforms in chemistry<br />

Formally, the transformation (Eq. 5.152) is a Laplace transformation (see Appendix<br />

E). A Laplace transform generally described the transformation from a microscopic<br />

(“microcanonical”) quantity such as ( ) to a macroscopic (“macrocanonical”, i.e.,<br />

dependent) quantity such as ( ).<br />

I Definition of the Laplace Transformation: The Laplace transform L [ ()] of a<br />

function () is defined as the integral<br />

() =L [ ()] =<br />

Z ∞<br />

0<br />

() − (5.153)<br />

I<br />

Inverse Laplace Transformation:<br />

with an arbitrary real number .<br />

() =L −1 [ ()] = 1<br />

2<br />

Z<br />

+∞<br />

−∞<br />

() (5.154)


5.4 Advanced collision theory 122<br />

I<br />

Notes:<br />

(1) The so far best known Laplace transform to us is the partition function<br />

( )= X <br />

− =<br />

Z ∞<br />

0<br />

() − (5.155)<br />

(2) While the pathway from ( ) to ( ) by forward Laplace transformation is<br />

straightforward, it is not generally straightforward to derive ( ) by inverse<br />

Laplace transformation from ( ), because ( ) would be needed from =0<br />

to = ∞.


5.4 Advanced collision theory 123<br />

5.4.5 Bimolecular rate constants from collision theory<br />

I Transformation from velocity to energy space: Taking our results from kinetic<br />

gas theory, we are now in the position to apply the fundamental equation of collision<br />

theory (Eq. 5.140)<br />

( )=<br />

Z ∞<br />

0<br />

() ( ) (5.156)<br />

using various functional forms for . Towards these ends, we first transform from to<br />

by inserting () from Eq. 5.64 (with the reduced mass instead of ). We thus<br />

obtain<br />

µ <br />

( )=4<br />

2 <br />

32<br />

∞<br />

As done above, we substitute for according to<br />

and obtain<br />

( )=<br />

Z<br />

0<br />

3 ()exp<br />

µ− 2 (5.157)<br />

2 <br />

= 1 2 2 (5.158)<br />

µ 12 2<br />

y =<br />

(5.159)<br />

<br />

y<br />

µ 12 2 1<br />

=<br />

2<br />

µ 1<br />

2<br />

y =<br />

µ 12 µ 1 2<br />

<br />

µ 12 µ 12 µ 12 1 1 1<br />

=<br />

(5.160)<br />

2 <br />

12 µ 1<br />

<br />

12<br />

(5.161)<br />

32<br />

∞<br />

Z<br />

0<br />

( )exp<br />

µ<br />

− <br />

<br />

<br />

<br />

(5.162)<br />

The integration from 0 to ∞ takes into account that a reaction can occur only if<br />

0 .<br />

This expression is interpreted as follows:<br />

• The function − describes the fraction of molecules with energy .<br />

• The function ( ) in the integrand is called the excitation function 38 . This<br />

can, in general, be a rather complicated (and not necessarily smooth) function<br />

(see Fig. 5.17).<br />

• The complete integrand ( ) − is called the reaction function 39 ,<br />

because it describes the reactive fraction of the molecules.<br />

• Therateconstant ( ) corresponds to the area under the reaction function.<br />

38 “Anregungsfunktion”<br />

39 “Reaktionsfunktion”


5.4 Advanced collision theory 124<br />

I<br />

Figure 5.17: Advanced collision theory: The excitation function.<br />

I<br />

Figure 5.18: Advanced collision theory: The reaction function.


5.4 Advanced collision theory 125<br />

I Non-equilibrium energy distribution:* The above expression for ( ) is valid if<br />

the reacting molecules are in internal equilibrium, i.e., the energy states are populated<br />

according to the Boltzmann distribution.<br />

Deviations from the Boltzmann distributions will occur in the case of very fast reactions,<br />

if the reaction is faster than the relaxation between the energy states. In this case, ( )<br />

decreases because the mean energy of the reacting molecule decreases.<br />

Mean energy of the reacting molecules (see Fig. 5.19):<br />

= X <br />

=<br />

Z ∞<br />

0<br />

() (5.163)<br />

(1) For thermal equilibrium:<br />

() =() (5.164)<br />

with () being the Boltzmann distribution.<br />

y<br />

(2) Non-equilibrium:<br />

() () (5.165)<br />

non-eq. eq. (5.166)<br />

I<br />

Figure 5.19: Mean energy of the reacting molecules in thermal equilibrium and for a<br />

non-eqilibrium situation.<br />

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


5.4 Advanced collision theory 126<br />

a) Advanced collision theory for hard spheres (Fig. 5.20)<br />

I<br />

Figure 5.20: Thehardspheremodel.<br />

(1) Ansatz for the reactive cross section:<br />

( )=<br />

(2) Inserting this ansatz into Eq. 5.162:<br />

( )=<br />

(3) Result for ( ): With<br />

we obtain<br />

y<br />

µ 12 µ 1 2<br />

<br />

Z<br />

½<br />

2 0<br />

0 0<br />

(5.167)<br />

32<br />

∞<br />

Z<br />

0<br />

2 exp<br />

µ<br />

− <br />

<br />

(5.168)<br />

<br />

= <br />

( − 1) (5.169)<br />

2 µ 12 µ 32 1 2<br />

( )=<br />

2<br />

<br />

µ<br />

× ( ) 2 × exp − µ<br />

<br />

− ∞<br />

<br />

<br />

¯¯¯¯ − 1 (5.170)<br />

0<br />

( )= 2 µ 8 <br />

<br />

12 µ<br />

1+ µ<br />

0<br />

exp − <br />

0<br />

<br />

(5.171)


5.4 Advanced collision theory 127<br />

(4) Arrhenius activation energy:<br />

y<br />

= 2 ln <br />

<br />

ln ∝ 1 µ<br />

2 ln +ln 1+ <br />

0<br />

− 0<br />

<br />

ln <br />

= 1 µ<br />

1<br />

2 + 1<br />

− µ <br />

0 0<br />

+<br />

1+ 0 2 2<br />

= 0 + 1 2 − 0<br />

1+ 0 <br />

(5.172)<br />

(5.173)<br />

(5.174)<br />

(5.175)<br />

b) Line-of-Centers model (Figs. 5.21 - 5.22)<br />

Even for hard spheres, because of angular momentum conservation (see below), not the<br />

entire collision energy, but only the fraction “directed” along the line of centers ()<br />

is effective.<br />

In particular, we have<br />

• for a collision with =0:<br />

= (5.176)<br />

y central collisions are most effective,<br />

• for a collision with = :<br />

=0 (5.177)<br />

y glancing collisions are ineffective because =0,<br />

• general expression (see below):<br />

= <br />

µ<br />

1 − 2<br />

2 <br />

(5.178)<br />

y decreases with increasing .


5.4 Advanced collision theory 128<br />

I<br />

Figure 5.21: Collision energy along line-of-centers<br />

Functional expression for ( ):<br />

(1) We assume that a reaction can occur only if ≥ 0 . Thus, we first determine<br />

(see Fig. 5.21):<br />

y<br />

= cos (5.179)<br />

2 = 2 cos 2 (5.180)<br />

= ¡ 2 1 − sin 2 ¢ (5.181)<br />

µ <br />

= 2 1 − 2<br />

(5.182)<br />

2<br />

= <br />

µ<br />

1 − 2<br />

2 <br />

(5.183)<br />

(2) We see from this expression that decreases with increasing . Thus, for a<br />

given ,wehaveamaximalvalueof (i.e., a value max ), for which ≥ 0 .<br />

We determine max from the condition that<br />

µ <br />

= 1 − 2<br />

≥ <br />

2 0 (5.184)<br />

y<br />

y<br />

2 ≤ 2 µ<br />

1 − 0<br />

<br />

<br />

= 2 max (5.185)<br />

( )= 2 max = 2 µ<br />

1 − 0<br />

<br />

<br />

(5.186)


5.4 Advanced collision theory 129<br />

y<br />

⎧<br />

⎨<br />

( )=<br />

⎩<br />

2 µ<br />

1 − 0<br />

<br />

<br />

0<br />

(5.187)<br />

0 0<br />

( ) is shown as function of in Fig. 5.22. Crossed molecular beam experiments<br />

have indeed shown similar dependencies for a number of reactions.<br />

I<br />

Figure 5.22: Reactive cross section for the line-of-centers model.<br />

(3) Result for ( ) from Eq. 5.162:<br />

y<br />

µ 1<br />

( )=<br />

<br />

Z ∞ µ<br />

× 2<br />

0<br />

=<br />

µ 1<br />

<br />

×<br />

Z ∞<br />

0<br />

( )=<br />

12 µ 2<br />

<br />

32<br />

1 − <br />

0<br />

<br />

12 µ 2<br />

32<br />

<br />

2 ( − 0 )exp<br />

µ 8 <br />

<br />

µ<br />

exp − <br />

<br />

(5.188)<br />

<br />

µ<br />

− <br />

<br />

(5.189)<br />

<br />

12 µ<br />

2 exp − <br />

0<br />

<br />

(5.190)<br />

This result is identical to that of our primitive “hard sphere-fixedmeanrelative<br />

speed” model from Section 5.1!


5.4 Advanced collision theory 130<br />

(4) Arrhenius activation energy:<br />

y<br />

= 2 ln <br />

<br />

(5.191)<br />

= 0 + 1 (5.192)<br />

2<br />

c) Allowance for the “steric effect”<br />

In order to account for the “structures” of the molecules which can undergo a reaction<br />

only if the reaction partners have a specificrelativeorientation,weintroducea(generally<br />

-dependent) steric factor ( ):<br />

µ 12 µ<br />

8 <br />

( )=( )<br />

2 exp − <br />

0<br />

(5.193)<br />

<br />

<br />

In ( ), we include all sorts of factors leading to deviations from the predicitions by simple<br />

collision theory. In favorable cases, ( ) can be predicted by theory. Towards these<br />

ends, we define a reaction probability ( ) (also called the opacity function,<br />

andcalculatedbytheory)andintegrateaccordingto<br />

( )=<br />

Z<br />

0 max<br />

( )2 (5.194)<br />

0<br />

In other cases, it may be convenient to define an angle dependent reaction threshold<br />

energy ∗ 0:<br />

∗ 0 = 0 + 0 (1 − cos ) (5.195)<br />

y<br />

∗ 0 = 0 for =0<br />

∗ (5.196)<br />

0 0 for 6= 0<br />

Such models are helpful for explaining steric factors in the range of 1 ≤ ≤ 0001.


5.4 Advanced collision theory 131<br />

I<br />

Figure 5.23: Collision theory: Allowance for steric effects.<br />

I<br />

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


5.4 Advanced collision theory 132<br />

I Harpooning: The value of ≥ 1 for the steric factor of the K + Br 2 reaction can<br />

be understood by the “harpooning mechanism”: An electron jumps from the K to the<br />

Br 2 at a long intermolecular distance. The subsequent collision and reaction at closer<br />

distance then occurs between a K + and Br − 2 . The collision between these two is strongly<br />

affectedbytheCoulombattraction.<br />

5.4.6 Long-range intermolecular interactions<br />

To this point, we have neglected any interactions between the reacting molecules other<br />

than the the collision itself. Thus, only straight line trajectories were allowed. In reality,<br />

however, there will be attractive or repulsive interactions between the colliding molecules<br />

(see Fig. 5.24). These will inevitably affect the collision dynamics. An extreme example<br />

is the K + Br 2 reaction mentioned above.<br />

We will investigate the potential energy hypersurfaces governing the reactions in Section<br />

6. Here, we will consider only long-range intermolecular interactions which can be easily<br />

implementedincollisiontheory.<br />

I<br />

Figure 5.24: Examples of attractive and repulsive trajectories for the collision between<br />

two molecules.<br />

a) Types of long-range interactions<br />

The long-range interactions between molecules can be described using multipole expansions<br />

of the potentials (charge: 1; dipole: 2; induced dipole: 3; quadrupole: 3 . . . In<br />

general, the interaction potential between two multipoles of ranks and 0 is proportional<br />

to − with = + 0 − 1:<br />

µ <br />

() =−<br />

(5.197)


5.4 Advanced collision theory 133<br />

I<br />

Table 5.3: Types of long-range intermolecular interactions.<br />

system<br />

()<br />

charge-charge (Coulomb) () = 1 2 2<br />

4 0 <br />

charge-dipole<br />

( ) = −→ · −→ = − cos <br />

4 0 2<br />

dipole-dipole ( 1 2 )=<br />

charge-quadrupole Θ ∝ −3<br />

dipole-quadrupole<br />

Θ ∝ −4<br />

charge-induced dipole ∝ −4<br />

quadrupole-quadrupole ΘΘ ∝ −5<br />

induced dipole-induced dipole <br />

∝ −6<br />

μ 1 · μ 2 − 3(μ 1 · r)(μ 2 · r)<br />

2<br />

4 0 3<br />

I Table 5.4: Long-range intermolecular interactions ( =15 D, =3Å 3 , =1,<br />

=0;energies in kJ mol; forcomparison: =25kJ mol @ =298K).<br />

³ ´ ³ ´<br />

system =5Å =10Å<br />

ion-ion 278 139<br />

ion-dipole 17.4 4.35<br />

ion-induced dipole 3.34 0.21<br />

dipole-dipole 2.17 0.27<br />

b) Lennard-Jones (6-12) potential for nonpolar molecules<br />

A realistic intermolecular potential for non-polar molecules is the Lennard-Jones (6-12)<br />

potential<br />

∙ ³LJ ´12 ³ LJ<br />

´6¸<br />

LJ () =4 LJ − (5.198)<br />

<br />

LJ is usually given in reduced form in units of K as<br />

∗ LJ = LJ<br />

<br />

(5.199)


5.4 Advanced collision theory 134<br />

I<br />

Figure 5.25: Schematic plot of the Lennard-Jones potential.<br />

I Estimation of LJ parameters: Values for LJ and LJ are typically derived from<br />

transport properties (e.g., viscosity), thermodynamic data (vdW parameters), or <strong>—</strong> if<br />

there is no better way <strong>—</strong> from empirical correlations (see, e.g., Svehla1963) with critical<br />

data, boiling temperatures, . . .<br />

I<br />

Table 5.5: Lennard-Jones for some gases (Mourits1977).<br />

LJ (pm) ( LJ ) (K) LJ (pm) ( LJ ) (K)<br />

He 2560 102 N 2 3738 820<br />

Ne 2822 320 O 2 3480 1026<br />

Ar 3465 1135 CH 4 3790 1421<br />

Kr 3662 1780 CF 4 4486 1673<br />

Xe 4050 2302 CCl 4 5611 4155<br />

H 2 2930 370 CO 2 3943 2009<br />

SF 6 5199 2120


5.4 Advanced collision theory 135<br />

I<br />

Figure 5.26: Lennard-Jones interaction potentials for He, Ne, Ar, Kr, and Xe.<br />

c) Orbital angular momentum and centrifugal barrier*<br />

In the case of a non-central collision, part of the translational energy is converted<br />

to “rotational energy” (orbital angular momentum; see Fig. 5.27). This is seen by<br />

partitioning the velocity vector between the direction ( )andtheradial<br />

direction ( ). The component corresponds to a rotational motion.<br />

I<br />

Figure 5.27: Orbital angular momentum in binary collisions.


5.4 Advanced collision theory 136<br />

• Radial velocity component:<br />

y<br />

• Kinetic energy:<br />

• Angular velocity:<br />

=<br />

cos = = <br />

<br />

= <br />

1<br />

2 2 +<br />

| {z }<br />

translation via LC<br />

1<br />

2 2 <br />

| {z }<br />

= 1 2 2 2 (rotation)<br />

(5.200)<br />

(5.201)<br />

(5.202)<br />

= <br />

= <br />

= <br />

= <br />

<br />

= <br />

2 (5.203)<br />

• Kinetic energy:<br />

= 1 2 ( ) 2 + 1 2 ( ) 2 (5.204)<br />

= 1 2 · 2 + 1 2<br />

2 2 2 (5.205)<br />

= 1 2 · 2 + 1 2 2<br />

<br />

2 (5.206)<br />

= 1 2 · 2 + 1 2 2 2<br />

2 2 (5.207)<br />

• Orbital angular momentum: is the orbital angular momentum <br />

= 2 <br />

= = (5.208)<br />

y<br />

= 1 2 · 2 +<br />

2<br />

2 2 (5.209)<br />

From quantum mechanics, 2 = ~ 2 ( +1)with as orbital angular momentum<br />

quantum number. y<br />

= 1 2 · 2 + ~2 ( +1)<br />

2 2 (5.210)<br />

= 1 2 · 2 + 2 ( +1) (5.211)<br />

where 2 = ~2<br />

is formally an -dependent rotational constant.<br />

22 • Centrifugal barrier: The centrifugal kinetic energy term ~2 ( +1)<br />

acts as a<br />

2 2<br />

centrifugal barrier<br />

centrifugal = 1 2 2<br />

<br />

= ~2 ( +1)<br />

(5.212)<br />

2 2 2


5.4 Advanced collision theory 137<br />

• Withalong-rangeattractivetermoftheform− ¡ <br />

¢ <br />

<br />

, the intermolecular potential<br />

thus has the form<br />

µ 2<br />

() =− + (5.213)<br />

<br />

µ 2 <br />

= − + ~2 ( +1)<br />

(5.214)<br />

2 2<br />

i.e.,<br />

() ∝− − + −2 (5.215)<br />

<strong>—</strong> The −2 term is important in practice if 3. In effect, it leads to an<br />

effective energy barrier (“centrifigual” barrier).<br />

<strong>—</strong> The larger for a given , the larger the orbital angular momentum .<br />

I<br />

Figure 5.28: Potential energy curves with centrifugal barriers for a diatomic system.<br />

d) Langevin “capture” rate constant for ion-molecule reactions*<br />

An nice illustration of the above results is the rate constant for ion-molecule reactions.<br />

The intermolecular interaction is governed by the point charge-induced dipole interaction.<br />

40 In addition, the centrifugal barrier has to be taken into account, so that<br />

() =−<br />

µ 2 ()2 <br />

(4 0 ) 2 2 + 4 (5.216)<br />

<br />

The resulting expression for the rate constant is<br />

à !<br />

4 () 2 12 µ 1<br />

( )=<br />

(4 0 ) 2 × Γ<br />

(5.217)<br />

2<br />

Typical values of are 100 × <br />

40 () needs to be checked again!!


5.4 Advanced collision theory 138<br />

6. Potential energy surfaces for chemical reactions<br />

6.1 Diatomic molecules<br />

I<br />

Morse potential (anharmonic oscillator):<br />

() = [1 − exp (− ( − ))] 2 (6.1)<br />

I<br />

Vibrational states of a harmonic oscillator:<br />

µ<br />

= ~ + 1 µ<br />

= + 1 <br />

2<br />

2<br />

(6.2)<br />

I<br />

Vibrational states of an anharmonic oscillator:<br />

µ<br />

= ~ + 1 µ<br />

− ~ + 1 2<br />

+ (6.3)<br />

2<br />

2<br />

I Dissociation energies D and D 0 :<br />

= 0 + 1 2 ~ − 1 4 ~ + (6.4)<br />

I<br />

Figure 6.1: Potential energy curve of diatomic molecules.


6.2 Polyatomic molecules 139<br />

6.2 Polyatomic molecules<br />

I<br />

I<br />

The potential energy becomes dependent of the different internal coordinates.<br />

Question: How many internal coordinates (dimensions) do we need?<br />

Answer:<br />

• Total number of degrees of freedom of an -atomic molecule:<br />

• for translational motion of center of gravity:<br />

• for rotational motion around center of gravity<br />

• resulting number of vibrational coordinates:<br />

3 (6.5)<br />

3 (6.6)<br />

2 for linear molecules (6.7)<br />

3 for nonlinear molecules (6.8)<br />

<strong>—</strong> linear molecules:<br />

<strong>—</strong> nonlinear molecules:<br />

3 − 5 (6.9)<br />

3 − 6 (6.10)<br />

I<br />

3N − 6 dimensional potential energy hypersurface:<br />

• The internal coordinates are described using the normal coordinates (vibrational<br />

displacements).<br />

• The potential energy function<br />

( 1 2 3−6 ) (6.11)<br />

is a hypersurface in a 3 − 6 (3 − 5) dimensional hyperspace. But we can plot<br />

only in 2-D or 3-D space.


6.2 Polyatomic molecules 140<br />

I<br />

Figure 6.2: Potential energy hypersurface for a simple atom transfer reaction (plot for<br />

fixed angle, e.g., collinear collision and other fixed coordinates depending on dimensionality<br />

of the system).<br />

I<br />

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


6.2 Polyatomic molecules 141<br />

I<br />

Figure 6.4: Potential energy hypersurface of HCO.<br />

I<br />

Figure 6.5: Potential energy hypersurface of HCO.


6.2 Polyatomic molecules 142<br />

I<br />

Models of potential energy surfaces:<br />

(1) Empirical or semiempirical models (e.g., LEPS),<br />

(2) Single points ( ) by ab initio calculations using quantum chemistry. Problem:<br />

Fitting the surface to the computed points.<br />

I<br />

<strong>Kinetics</strong> and Dynamics:<br />

(1) TST = Transition State Theory (statistical theory),<br />

(2) Classical trajectory calculations (by solving Newton’s equations),<br />

(3) Fully quantum mechanical wave packet calculations (by solving the timedependent<br />

Schrödinger equation).


6.3 Trajectory Calculations 143<br />

6.3 Trajectory Calculations<br />

The “exact” dynamics of the molecules on a potential energy hypersurface can be<br />

followed by classical trajectory calculations.<br />

I Phase Space: 6-dimensional space spanned by the spatial coordinates ( )and<br />

the conjugated momenta ( ):<br />

↔ (6.12)<br />

where<br />

= <br />

<br />

<br />

= <br />

·<br />

(6.13)<br />

I<br />

I<br />

Hamilton function:<br />

= + = 2<br />

+ () (6.14)<br />

2<br />

Note that depends only on and depends only on .<br />

Equations of motion:<br />

• for a 1 particle system:<br />

(1)<br />

= = <br />

<br />

= −<br />

()<br />

<br />

y<br />

·<br />

= − <br />

= − <br />

since contains only not .<br />

(2)<br />

y<br />

y<br />

• for an atomic system:<br />

= + =<br />

(6.15)<br />

(6.16)<br />

= 2<br />

2 + (6.17)<br />

<br />

<br />

= 1 × = 1 µ<br />

× <br />

·<br />

3X<br />

=1<br />

·<br />

= <br />

<br />

= ·<br />

(6.18)<br />

(6.19)<br />

<br />

2 2 + ( 1 2 3−6 ) (6.20)<br />

·<br />

= − <br />

(6.21)<br />

<br />

·<br />

= <br />

(6.22)<br />

<br />

Thus we obtain a set of coupled differential equations which can be solved using<br />

the Runge-Kutta or Gear methods.


6.3 Trajectory Calculations 144<br />

• Thermal rate constants ( ) follow by suitable averaging over the initial conditions<br />

( vibrational states, rotational states, angles, ...).<br />

• Major problems: Zero-point energy? Tunneling? Surface crossings?<br />

I<br />

Figure 6.6: Trajectory calculations.<br />

I<br />

Figure 6.7: Trajectory calculations.


6.3 Trajectory Calculations 145<br />

I<br />

Figure 6.8: The reaction H + Cl 2 isareactionwithanearlyTS.Thereactionisknown<br />

to yield HCl with an inverted vibrational state distribution (⇒ HCl-Laser).<br />

I<br />

Figure 6.9: Bimolecular reaction with a late TS.


6.3 Trajectory Calculations 146<br />

7. Transition state theory<br />

7.1 Foundations of transition state theory<br />

The concept of transition state theory (TST) or activated complex theory (ACT) goes<br />

back to H. Eyring and M. Polanyi (1935). TST is a statistical theory; it does not give<br />

information on the exact dynamics of a reaction. It does, however, give (statistical)<br />

information on product state distributions.<br />

There are two forms of TST:<br />

• Macrocanonical TST ⇒ ( ) for a macrocanonical ensemble of molecules with<br />

aconstant (an ensemble of molecules with energy states populated according<br />

to the Boltzmann distribution - each state being equally likely.)<br />

• Microcanonical TST (TST) ⇒ () for a microcanonical ensemble of molecules<br />

with specific energy.<br />

We will consider only the first version.<br />

The detailed application of TST also depends on whether or not the reaction has a<br />

distinctive potential energy barrier (⇒ conventional TST or variational TST).<br />

I<br />

Figure 7.1: Transition State Theory (TST): Fundamental equations and applications.


7.1 Foundations of transition state theory 147<br />

7.1.1 Basic assumptions of TST<br />

(1) The Born-Oppenheimer approximation (separation of electronic and nuclear motion)<br />

is a valid approximation. Thus we can describe the nuclear motion during<br />

the reaction on a single electronic potential energy surface (PES; see Fig. 7.2). 41<br />

(2) The reactant molecules are in internal equilibrium (the quantum states are populated<br />

according to the Boltzmann distribution).<br />

(3) “Transition state” and “dividing surface“:<br />

a) Conventional TST (valid for so-called “type I” PES’s = PES’s with a distinctive<br />

energy barrier along the reaction coordinate (RC)): We can localize<br />

a distinctive “transition state” (TS) for the reaction at the saddle point on<br />

the PES that separates the reactant and product “valleys”. The TS is located<br />

at the maximum of the PES along the minimum energy path (MEP;<br />

see Fig. 7.3).<br />

b) Variational TST (version that can be used for so-called “type II” PES’s =<br />

PES’s without a clear energy barrier along the RC (we can’t localize a simple<br />

TS)): We can localize a 3 − 7-dimensional “dividing surface” (DS) on the<br />

3 − 6-dimensional PES that separates the reactant and product regions<br />

on the PES in such a way that the “reactive flux” through the DS becomes<br />

minimal (Fig. 7.2). The TS is then placed at the point where the MEP<br />

intersects the DS. In this case, the position of the TS depends strongly on<br />

the finer details such as angular momentum (because of centrifugal barriers<br />

which move inwards with increasing or increasing ).<br />

(4) The motion along the minimum energy path across the TS (through the DS) can<br />

be separated from other degrees of freedom and treated as translational motion.<br />

AttheTS,wethereforehave1translationaldegreeoffreedom ‡ (the RC) and<br />

3 − 7 (vibrational) degrees of freedom orthogonal to the RC.<br />

(5) Reactant molecules passing through the DS never come back (no recrossing).<br />

(6) The states of the TS are in quasi-equilibrium with the states of the reactant<br />

molecule even if we have no equilibrium between reactants and products (the<br />

individual molecules don’t know that this equilibrium is not present - this quasiequilibrium<br />

assumption can be dropped in dynamical derivations of TST; see<br />

below).<br />

I<br />

Conditions for TS at the saddle point:<br />

(1) has maximum along ‡ :<br />

<br />

=0and 2 <br />

0 (7.1)<br />

‡ <br />

‡2<br />

y 1 imaginary frequency (criterion for TS in quantum chemical calculations)!<br />

41 TST cannot be used at least in the simple form considered in this lecture when two potential surfaces<br />

are involved that are energetically close to each other (or even become degenerate) and the reaction<br />

involves a “non-adiabatic transition” between these two potentials.


7.1 Foundations of transition state theory 148<br />

(2) has minimum along all 3 − 7 other coordinates:<br />

<br />

<br />

=0and 2 <br />

2 <br />

0 (7.2)<br />

y 3 − 7 real other frequencies.<br />

I<br />

Figure 7.2: Transition state theory: The dividing surface with minimal reactive flux.<br />

I<br />

Figure 7.3: Minimum energy path (MEP) and transition state (TS) of a simple atom<br />

transfer reaction.


7.1 Foundations of transition state theory 149<br />

7.1.2 Formal kinetic model<br />

In order to derive the TST expression for ( ), wefirst look at a simple formal kinetic<br />

model. This formal kinetic model has its roots in the original idea that the TS can be<br />

considered as some sort of an “activated complex” with a local energy minimum on<br />

the PES. It is not a good derivation of the TST expression for ( ), but leads to the<br />

correct expression.<br />

I<br />

Activated complex model of TST:<br />

y<br />

A+BC 1 À<br />

2<br />

ABC ‡ 3 → P (7.3)<br />

£<br />

ABC<br />

‡ ¤ = 1 [A] [BC]<br />

2 + 3<br />

(7.4)<br />

I Assumption: 3 ¿ 2 y<br />

£ ¤ ABC<br />

‡ = 1<br />

[A] [BC] = <br />

12 [A] [BC] (7.5)<br />

2<br />

y Equilibrium between the reactants (A+BC)andtheTS(ABC ‡ ):<br />

y<br />

[P]<br />

<br />

= 3<br />

£<br />

ABC<br />

‡ ¤ <br />

(7.6)<br />

= 3 | {z 12 [A] [BC] (7.7)<br />

}<br />

=<br />

I<br />

ACT expression for rate constant:<br />

= 3 12 (7.8)<br />

Below, we will find that<br />

and<br />

3 = ‡ = <br />

<br />

<br />

12 ∝ exp<br />

µ− ∆‡<br />

<br />

(7.9)<br />

(7.10)<br />

7.1.3 The fundamental equation for k (T )<br />

The above derivation of the TST rate constant was based on an oversimplified picture,<br />

since the TS is not at a local minimum but at a saddle point on the PES. In this section,<br />

we derive the TST fundamental equation more carefully by bringing in some dynamical<br />

aspects. In (1) - (4), we explore the implications of the quasi-equilibrium assumption,<br />

and in (5) - (9) we develop a dynamical semiclassical model: 42<br />

42 Semiclassical because the translation is handled classically while all other degrees of freedom are<br />

described quantum mechanically.


7.1 Foundations of transition state theory 150<br />

(1) We start by assuming equilibrium between the reactants and products:<br />

A+BC 1 À<br />

2<br />

ABC ‡ 3<br />

À<br />

4<br />

P (7.11)<br />

(2) We consider two dividing surfaces at the TS separated by a distance of ≤ .<br />

is the “length” of a “phase space cell” in semiclassical theory which is determined<br />

by Heisenberg’s uncertainty principle:<br />

∆ ‡ × ∆ ‡ = × ∆ ‡ = (7.12)<br />

(3) We now consider the number densities of molecules passing through these surfaces<br />

from left to right ( −→ ‡ )andfromrighttoleft( ←−<br />

‡ ). The total number density ‡<br />

at the TS in equilibrium is their sum, i.e.,<br />

‡ = −→ ‡ + ←−<br />

‡ = ‡ A BC (7.13)<br />

In equilibrium, we must have<br />

Thus,<br />

−→<br />

‡ = ←−<br />

‡ (7.14)<br />

−→<br />

‡ = ←−<br />

‡ = 1 2 ‡ = ‡ <br />

2 A BC (7.15)<br />

(4) We now suddenly remove all the products. Then ←−<br />

‡ =0 However, there is no<br />

reason for −→ ‡ to change, because the molecules on the reactant side remain in<br />

internal equilibrium (this is the quasi-equilibrium assumption). 43 y<br />

−→<br />

‡ = ‡<br />

2 = ‡ <br />

2 A BC (7.16)<br />

Thus, we can still express −→ ‡ using the equilibrium constant ‡ even if there is<br />

no equilibrium between reactants and products.<br />

(5) We now turn to the reaction rate, which is given by the number density of molecules<br />

−→ ‡ in the time interval that are crossing the DS in the forward direction.<br />

We can write this as<br />

−→ ‡<br />

= −→ −→<br />

<br />

‡ ‡ ‡<br />

=<br />

<br />

‡ (7.17)<br />

where<br />

• −→ ‡ is the decay rate of the TS in the forward direction (or the frequency<br />

factor for the molecules crossing the DS in the forward direction),<br />

43 The quasi-equilibrium assumption for −→ ‡ is made in canonical TST. In a microcanonical derivation<br />

· ·<br />

−→ ←−<br />

of the TST expression for (), we consider the fluxes ‡ and ‡ through the dividing surface<br />

at the TS. Then, the individual molecules “dont’t know” that the products have been removed and<br />

·<br />

·<br />

←−<br />

−→<br />

‡ =0. Since they don’t know about this, the flux ‡ has to remain unchanged.


7.1 Foundations of transition state theory 151<br />

• −→ ‡ = −→ ‡ is the mean speed of the molecules crossing the DS in the forward<br />

direction.<br />

(6) We now use the equilibrium constant to determine ‡ via<br />

‡ = ‡ ABC<br />

A BC<br />

(7.18)<br />

y ‡ = ‡ ABC = ‡ A BC (7.19)<br />

y<br />

−→ −→<br />

‡ <br />

‡<br />

=<br />

‡ A BC = A BC (7.20)<br />

with<br />

−→<br />

<br />

‡<br />

=<br />

‡ (7.21)<br />

(7) From statistical thermodynamics (see Appendix G), we have the following expression<br />

for the equilibrium constant ‡ ( ) in terms of the molecular partition<br />

functions:<br />

‡ ( )=<br />

∗ <br />

=<br />

µ<br />

<br />

exp − ∆ <br />

0<br />

(7.22)<br />

<br />

• The ’s are the respective molecular partition functions (per unit volume).<br />

• ∗ is written with respect to the zero-point level of the reactants.<br />

• The exp (−∆ 0 ) factor arises subsequently, because we now express<br />

the value of partition function for the TS relative to the zero-point level of<br />

the TS ( ∗ = exp (−∆ 0 )).<br />

(8) Separating the motion along the reaction coordinate ‡ from the other degrees of<br />

freedom, we write<br />

= ‡ ‡ <br />

(7.23)<br />

where ‡ is the 1-D (translational) partition function describing the translational<br />

motion of the molecules along ‡ across the TS and ‡ <br />

is the partition function<br />

for all 3 − 7 other (rotational-vibrational-electronic) degrees of freedom.<br />

Here,<br />

• the 1-D translational partition function is<br />

‡ =<br />

µ 2 <br />

2 12<br />

× (7.24)<br />

• the mean speed along ‡ for crossing the DS in the forward direction is given<br />

by<br />

R<br />

−→ ∞<br />

‡ −2 2 µ <br />

<br />

12<br />

0<br />

= R ∞<br />

−∞ −2 2 = <br />

(7.25)<br />

2


7.1 Foundations of transition state theory 152<br />

(9) Collecting and inserting the results into<br />

−→<br />

<br />

‡<br />

( )=<br />

( ) (7.26)<br />

we obtain<br />

( )= 1 µ 12 µ 12 µ<br />

2 <br />

×<br />

× ‡ <br />

exp − ∆ <br />

0<br />

(7.27)<br />

2<br />

2<br />

<br />

y<br />

( )= ‡ µ<br />

<br />

exp − ∆ <br />

0<br />

(7.28)<br />

<br />

I Result: In order to account () forsomerecrossingoftheTSand() for quantum<br />

mechanical tunneling, we add an additional factor called the transmission coefficient.<br />

For well-behaved reactions, ≈ 1. Thus, we end up with the fundamental equation<br />

of TST in the form:<br />

( )= <br />

<br />

‡ µ<br />

<br />

exp − ∆ <br />

0<br />

<br />

Note that ‡ <br />

is the partition function of the TS excluding the RC.<br />

Equation 7.29 allows us to calculate ( ) using the following data:<br />

(7.29)<br />

• The threshold energy for the reaction ∆ 0 (or, per mol, ∆ 0 ). ∆ 0 has to be<br />

obtained from ab initio quantum chemistry calculations, and<br />

• structural information on the reactants and the TS (i.e., bond lengths and angles,<br />

vibrational frequencies).<br />

I Figure 7.4: Illustration of ∆ ‡ , ∆ 0 ,and .


7.1 Foundations of transition state theory 153<br />

∆ ‡ : Born-Oppenheimer potential energy barrier (7.30)<br />

∆ 0 : zero-point corrected threshold energy for the reaction (7.31)<br />

: Arrhenius activation energy (7.32)<br />

∆ 0 is given with respect to the zero-point energies of the reactants and the TS,<br />

= 1 2<br />

X<br />

(7.33)<br />

7.1.4 Tolman’s interpretation of the Arrhenius activation energy<br />

I We now understand the difference between E and ∆E 0 :<br />

2 ln ( )<br />

= <br />

<br />

Ã<br />

= 2 <br />

ln <br />

<br />

<br />

‡ µ<br />

<br />

exp<br />

<br />

= ∆ 0 + + 2 ln ‡ <br />

−<br />

| {z }<br />

=h reacting moleculesi<br />

− ∆ 0<br />

<br />

(7.34)<br />

! (7.35)<br />

µ<br />

2 ln <br />

+ 2 ln <br />

<br />

(7.36)<br />

<br />

<br />

| {z }<br />

=h all molecules i<br />

• ∆ 0 is difference of the zero-point levels of the reactants and the TS,<br />

• the term is the mean translational energy in the RC (from the factor),<br />

• the 2 ln terms are the internal (vibration-rotation) energies of the TS and<br />

<br />

the reactants, respectively. y<br />

I Tolman’s interpretation of E (see Fig. 7.4):<br />

= h reacting molecules i − h all molecules i (7.37)


7.2 Applications of transition state theory 154<br />

7.2 Applications of transition state theory<br />

( )= <br />

<br />

‡ µ<br />

<br />

exp − ∆ <br />

0<br />

<br />

(7.38)<br />

TST has gained enormous importance in all areas of physical chemistry because it<br />

provides the basis for understanding – and/or calculating –<br />

• quantitative values of preexponential factors,<br />

• ∆ 0 from measured values,<br />

• -dependence of and deviations from Arrhenius behavior,<br />

• gas phase and liquid phase reactions (proton transfer, electron transfer, organic<br />

reactions, reactions of ions in solutions),<br />

• pressure dependence of reaction rate constants (activation volumes),<br />

• kinetic isotope effects,<br />

• structure-reactivity relations (e.g., Hammett relations in organic chemistry),<br />

• features of biological reactions (enzyme reactions),<br />

• heterogeneous reactions (heterogeneous catalysis, electrode kinetics), . . .<br />

7.2.1 The molecular partition functions<br />

In order to evaluate Eq. 7.29 for a given reaction, we need, besides ∆ 0 , the molecular<br />

partition functions for the reactants and the TS. We summarize only the main points<br />

here; further information is given in Appendix G.<br />

I<br />

Definition:<br />

= P − <br />

(7.39)<br />

I<br />

<strong>Physical</strong> interpretation:<br />

• The molecular partition function is a number which describes how many states<br />

are available to the molecule at a given temperature .<br />

• The ratio<br />

‡ <br />

(7.40)<br />

<br />

in Eq. 7.29 is thus the ratio of the number of states available to the TS and<br />

‡ <br />

the reactant molecules. Since each state is equally likely, is the relative<br />

<br />

probability of the TS vs. the probability of the reactants.<br />

Note again that ‡ <br />

motion along the RC.<br />

is the partition function of the TS excluding the translational


7.2 Applications of transition state theory 155<br />

I Separation of degrees of freedom: For separable degrees of freedom, we can write<br />

the molecular partition function as<br />

= × × × (7.41)<br />

Nuclear spin ( ) may have to be taken into account in some cases as well.<br />

I<br />

Table 7.1: Molecular partition functions.<br />

type of motion energy partition function magnitude<br />

µ <br />

translation (1-D) 2 2<br />

12 2 <br />

10 8 × <br />

8 2<br />

<br />

µ µ 2 <br />

translation (3-D) 2 <br />

2 32<br />

<br />

+ 2 <br />

+ 2 2 <br />

10 24 × <br />

8 2 2 2 <br />

µ 2 <br />

rotation (1-D) a ~ 2<br />

12<br />

2 2<br />

2 <br />

10<br />

µ ~ 2 <br />

rotation (2-D) b ~ 2<br />

2 <br />

( +1)<br />

10 2<br />

2 ~ 2 <br />

µ <br />

rotation (3-D) c 12<br />

32 2 <br />

or <br />

(<br />

~ 2 ) 12 10 3 10 4<br />

vibration d [1 − exp (− )] −1 1 10<br />

electronic <br />

P exp (− ) 1<br />

a Free internal rotor. For a hindered rotor (torsional vibrational mode), the partition function has to<br />

be determined by an explicit calculation over the hindered rotor quantum states.<br />

b Linear molecule.<br />

c Nonlinear molecule.<br />

d For each harmonic oscillator degree of freedom measured from zero-point level.


7.2 Applications of transition state theory 156<br />

7.2.2 Example 1: Reactions of two atoms A + B → A···B ‡ → AB<br />

This example is somewhat artifical unless there is some mechanism to remove the bond<br />

energy (for example, in associative ionization, A + B → A···B ‡ → AB + + − ).<br />

‡ µ<br />

<br />

exp − ∆ <br />

0<br />

<br />

Ã<br />

! Ã !<br />

= ‡ <br />

‡ <br />

( )( ) 1<br />

<br />

= µ<br />

2 ( + ) 2<br />

<br />

2 2 <br />

µ<br />

× exp − ∆ <br />

0<br />

<br />

= µ<br />

<br />

2<br />

+ <br />

2 <br />

µ<br />

× exp − ∆ <br />

0<br />

<br />

( )= <br />

<br />

<br />

µ<br />

exp − ∆ <br />

0<br />

<br />

2<br />

2 <br />

32 µ 8 2 ‡ <br />

2<br />

32<br />

Ã<br />

!<br />

8 2 ( + ) 2<br />

2 + <br />

y<br />

µ 12 µ<br />

8 <br />

( )=<br />

( + ) 2 exp − ∆ <br />

0<br />

<br />

<br />

This result is identical to that obtained by the LC model in collision theory!<br />

(7.42)<br />

(7.43)<br />

(7.44)<br />

(7.45)<br />

(7.46)


7.2 Applications of transition state theory 157<br />

7.2.3 Example 2: The reaction F + H 2 → F···H···H ‡ → FH + H<br />

I<br />

Exercise 7.1: Evaluatetherateconstantforthereaction<br />

F+H 2 → F ···H ···H ‡ → FH + H (7.47)<br />

using the data in Table 7.2 at =200, 300, 500, 1000 and 2000 K, determinethe<br />

Arrhenius parameters, and compare the results with the experimental value of<br />

=2× 10 14 exp (−800 )cm 3 mol s (7.48)<br />

¤<br />

I Table 7.2: Properties of the reactants and transition state of the reaction F + H 2 .<br />

parameter F ···H ···H ‡ a F H 2<br />

2 (F ···H) ( Å) 1602<br />

1 (H ···H) ( Å) 0756<br />

07417 b<br />

1 (cm −1 ) 40076 43952<br />

2 (cm −1 ) 3979<br />

3 (cm −1 ) 3979<br />

4 (cm −1 ) 3108 c<br />

³ (u) 21014 189984 2016<br />

u Å 7433<br />

0277<br />

1 2<br />

<br />

d<br />

4 4 1<br />

0 ¡ kJ mol −1¢ 657 000<br />

a The transition state FHH ‡ is assumed to be linear.<br />

b = (H-H).<br />

c One imaginary frequency describes the TS along the RC.<br />

d The electronic ground state of F is 2 32 . We neglect the 2 12 spin-orbit component at ( 2 12 )=<br />

404 cm −1 because it does not correlate with the products H + HF. The electronic state of linear<br />

FHH ‡ is 2 Π. The ground state of H 2 is 2 Σ + .<br />

I Solution 7.1:<br />

( )= ‡ µ<br />

<br />

exp − ∆ <br />

0<br />

2 <br />

Ã<br />

!<br />

= ‡ <br />

( )( 2 )<br />

µ<br />

× exp − ∆ <br />

0<br />

<br />

<br />

à !<br />

‡ <br />

2<br />

<br />

à !<br />

‡ <br />

2<br />

<br />

(7.49)<br />

à !<br />

‡ <br />

2<br />

<br />

(7.50)


7.2 Applications of transition state theory 158<br />

The different factors can be evaluated separately:<br />

Ã<br />

!<br />

‡ <br />

µ 32 µ <br />

+ 2 <br />

2 32<br />

=<br />

(7.51)<br />

( )( 2 )<br />

2 2 <br />

<br />

à !<br />

‡ µ <br />

<br />

8 2 ‡ 2<br />

‡<br />

=<br />

= <br />

2 8 2 2 2 2 2 (7.52)<br />

2<br />

<br />

à !<br />

‡ <br />

1 − exp (− 2 )<br />

=<br />

(7.53)<br />

2<br />

3Y<br />

´i<br />

<br />

à !<br />

‡ <br />

2<br />

<br />

=1<br />

= 4<br />

4 × 1<br />

h<br />

1 − exp<br />

³<br />

− ‡ <br />

(7.54)<br />

Note: The difference between the TST result and experiment can be ascribed (a)<br />

to tunneling and (b) deficiencies of the ab initio PES. Indeed, according to newer ab<br />

initio calculations, the TS is actually slightly bent and the TS parameters differ slightly,<br />

yielding better agreement with experiment. ¥


7.2 Applications of transition state theory 159<br />

7.2.4 Example 3: Deviations from Arrhenius behavior (T dependence of k (T ))*<br />

I<br />

Exercise 7.2: The temperature dependence of bimolecular gas phase reactions of the<br />

type A + B → productscanusuallybeexpressedintheform<br />

Using the TST expression for ( )<br />

( )= × × exp (−∆ 0 ) (7.55)<br />

( )= <br />

<br />

‡ µ<br />

<br />

exp<br />

<br />

− ∆ 0<br />

<br />

<br />

, (7.56)<br />

determine the values of for the types of reactions in Table 7.3, assuming that the<br />

vibrational degrees of freedom are not excited. ¤<br />

I Table 7.3:<br />

A + B → TS ‡ <br />

transl. rot. <br />

<br />

atom atom linear TS 1 32 1<br />

+05<br />

32 32 1<br />

atom linear molecule linear TS 1 −32 1<br />

−05<br />

1<br />

atom linear molecule nonlin. TS 1 −32 32<br />

00<br />

1<br />

atom nonlin. molecule nonlin. TS 1 −32 32<br />

−05<br />

32<br />

linear molecule linear molecule linear TS 1 −32 1<br />

−15<br />

1 1<br />

linear molecule linear molecule nonlin. TS 1 −32 32<br />

−10<br />

1 1<br />

linear molecule nonlin. molecule nonlin. TS 1 −32 32<br />

−15<br />

1 32<br />

nonlin. molecule nonlin. molecule nonlin. TS 1 −32 32<br />

−20<br />

32 <br />

32<br />

I Solution 7.2: See Table 7.3. ¥<br />

Using the above results, we can recast ( ) and determine the following expression for<br />

the Arrhenius activation energy, with as above:<br />

= ∆ 0 + + 2 ln ‡ <br />

<br />

− 2<br />

X<br />

Reactants<br />

ln reactants<br />

<br />

<br />

(7.57)<br />

We see that the value for depends on the vibrational frequencies of the TS and the<br />

reactants. Often, the frequencies of the reactants are very high so that reactant<br />

=1.<br />

However, the TS may have soft vibrations so that ‡ <br />

may approach the classical value<br />

=[1− exp (− )] −1 → <br />

for →∞ (7.58)<br />

<br />

Thus, each “soft” vibrational mode in the TS increases by 1, each “soft” vibrational<br />

mode in the reactants decreases by 1.


7.3 Thermodynamic interpretation of transition state theory 160<br />

7.3 Thermodynamic interpretation of transition state theory<br />

7.3.1 Fundamental equation of thermodynamic transition state theory<br />

As shown above,<br />

( )= <br />

‡ (7.59)<br />

<br />

‡ can often be interpreted using thermodynamic arguments, i.e., ∆ ‡ , ∆ ‡ ,and<br />

∆ ‡ . For liquid phase reactions, this is straightforward. For gas phase reactions, we<br />

have to take into account the difference between ‡ and .<br />

‡<br />

a) Thermodynamic transition state theory for unimolecular gas phase reactions<br />

For unimolecular isomerization and dissociation reactions of the type<br />

or<br />

we have<br />

ABC → ABC ‡ → ACB ‡ (7.60)<br />

ABC → ABC ‡ → A+BC ‡ , (7.61)<br />

£ ¤<br />

ABC<br />

‡<br />

‡ ( )=<br />

[ABC]<br />

= ABC ‡<br />

ABC<br />

= ‡ ( ) (7.62)<br />

‡ ( )= ‡ ( )= ‡ ( ) is dimensionless (∆ ‡ =0) and we do not have to worry<br />

about a difference between ‡ ( )= ‡ ( ) and about units or different standard<br />

states.<br />

I Results: Eq. 7.59 thus gives the following results:<br />

• Rate constant:<br />

y<br />

( )= <br />

‡ ( )= µ <br />

<br />

exp − ∆‡<br />

<br />

( )= µ <br />

<br />

∆<br />

‡<br />

exp exp<br />

µ− ∆‡<br />

<br />

<br />

(7.63)<br />

(7.64)<br />

• Arrhenius activation energy:<br />

y<br />

= 2 ln <br />

<br />

= 2 ∙ µ<br />

ln ∆<br />

‡<br />

exp <br />

<br />

exp<br />

¸<br />

µ− ∆‡<br />

<br />

(7.65)<br />

<br />

= ∆ ‡ + + 2 ∆‡<br />

<br />

= ∆ ‡ + + ∆‡ <br />

<br />

(7.66)<br />

(7.67)<br />

If ∆ ‡ 6= ( ) in the temperature range of interest, we obtain<br />

= ∆ ‡ + (7.68)


7.3 Thermodynamic interpretation of transition state theory 161<br />

• Arrhenius preexponential factor:<br />

∆ ‡ = − (7.69)<br />

y<br />

y<br />

( )= µ µ<br />

∆<br />

‡<br />

exp exp − <br />

<br />

<br />

= µ <br />

∆<br />

‡<br />

exp <br />

(7.70)<br />

(7.71)<br />

b) Thermodynamic transition state theory for bimolecular gas phase reactions<br />

I Conversion from K<br />

‡ (T ) to K‡ <br />

(T ): For bimolecular reactions<br />

A+BC→ ABC ‡ , (7.72)<br />

we have to account for the different numbers of species ∆ ‡ = ‡ − reactants :<br />

We therefore have to convert from ‡ ( ) to ‡ ( ) via<br />

∆ ‡ = −1 . (7.73)<br />

£ ¤ <br />

ABC<br />

‡ ABC<br />

‡ ABC ‡ ª<br />

‡ ( )=<br />

[A] [BC] = <br />

A BC<br />

= <br />

A ª BC ª × 1<br />

µ −∆<br />

‡<br />

<br />

= ª ‡ ( ) ×<br />

.<br />

ª<br />

<br />

(7.74)<br />

Here, ‡ ( ) is conveniently given in units of lmol −1 . 44 However, , ‡ ∆ ‡ , ∆ ‡ and<br />

∆ ‡ arerelatedtothestandardstateat ª =1bar. 45 To keep track of units in the<br />

conversion , it is useful to apply in<br />

lbar by writing<br />

mol K<br />

=83145<br />

J<br />

mol K =83145 m3 Pa<br />

lbar<br />

=0083145<br />

mol K mol K = 0 (7.76)<br />

44 With ‡ ( ) in units of ¡ mol l −1¢ ∆<br />

, we get as close as possible to the standard state ∗ =<br />

1molkg −1 for reactions in aqueous solution.<br />

45 We recall that the thermodynamic equilibrium constant ( )=exp(−∆ ª ) referred to<br />

the standarrd state at ª =1baris dimensionless. Instead of the dimensionless ‡ ( ), onemay<br />

also decide to use the equilibrium constant in pressure units ‡ 0 :<br />

hABC ‡i ABC ‡<br />

‡ ( )=<br />

[A] [BC] = <br />

A BC<br />

= ‡ 0 ( ) × 0 (7.75)


7.3 Thermodynamic interpretation of transition state theory 162<br />

so that, with ∆ ‡ = −1, weobtain<br />

Thus,viaEq.7.59:<br />

‡ ( )=<br />

( )= <br />

<br />

ABC ‡ ª<br />

A ª<br />

<br />

<br />

BC ª<br />

<br />

× 1<br />

µ 0<br />

= <br />

ª ‡ ( ) × . (7.77)<br />

ª<br />

0 <br />

ª ‡ ( )= µ<br />

0 <br />

exp ª<br />

− ∆ª‡<br />

<br />

<br />

, (7.78)<br />

in units of<br />

µ µ <br />

0 <br />

dim ( ( )) = dim × dim = l<br />

(7.79)<br />

<br />

ª mol s<br />

and with the standard state for the thermodynamics quantities explicitly indicated by<br />

the ª symbol.<br />

I<br />

Results:<br />

• Rate constant:<br />

( )= <br />

<br />

µ <br />

0 ∆<br />

ª‡<br />

exp exp<br />

ª <br />

µ− ∆ª‡<br />

<br />

<br />

(7.80)<br />

• Arrhenius activation energy:<br />

= 2 ln <br />

<br />

(7.81)<br />

If ∆ ª‡ 6= ( ) in the temperature range of interest:<br />

= ∆ ª‡ +2 (7.82)<br />

• Arrhenius preexponential factor:<br />

∆ ª‡ = − 2 (7.83)<br />

y<br />

( )= <br />

<br />

= 2 <br />

<br />

µ <br />

0 ∆<br />

ª‡<br />

exp ª <br />

µ<br />

0 ∆<br />

ª‡<br />

exp ª <br />

µ<br />

exp<br />

<br />

exp<br />

− <br />

− 2<br />

<br />

µ<br />

− <br />

<br />

<br />

(7.84)<br />

(7.85)<br />

Comparison with the Arrhenius expression = exp (− ) thus gives<br />

= 2 <br />

<br />

µ <br />

0 ∆<br />

ª‡<br />

exp ª <br />

(7.86)


7.3 Thermodynamic interpretation of transition state theory 163<br />

• Note: and are given above in units of lmol −1 s −1 (we may denote this by<br />

writing them as and ), while ∆ ª‡ and ∆ ª‡ used above are referred to<br />

the standard state for gases of ª =1bar.<br />

<strong>—</strong> For ideal gases, ∆ ‡ and ∆ ‡ are related via<br />

∆ ª‡ = ∆ ª‡ + ∆ ‡ ( )=∆ ª‡ − = ∆ +‡ − (7.87)<br />

<strong>—</strong> Likewise, ∆ ª‡ and ∆ +‡ for are related via<br />

∆ +‡ − ∆ ª‡ =<br />

y<br />

Z <br />

+<br />

ª<br />

Z<br />

<br />

+<br />

=<br />

ª<br />

<br />

= Z <br />

+<br />

ª<br />

<br />

= Z <br />

+<br />

ª<br />

∆ ‡ <br />

(7.88)<br />

<br />

= ∆ ‡ +<br />

ln<br />

+ 0 = +<br />

ª ∆‡ ln<br />

+ 0 = ª<br />

ª ∆‡ ln (7.89)<br />

+ 0 <br />

= ∆ ‡ ln + 0 <br />

(7.90)<br />

ª<br />

with ª =1barand + = +<br />

+ =1moll−1 .<br />

∆ ª‡ = ∆ +‡ − ln + 0 <br />

ª (7.91)<br />

c) Thermodynamic transition state theory for reactions in the liquid phase<br />

For reactions in condensed phases,<br />

∆ ≈ ∆ (7.92)<br />

and<br />

y<br />

£<br />

ABC<br />

‡ ¤<br />

‡ ( ) =<br />

[A] [BC] = ¡ +¢ ∆ ‡ exp<br />

≈ ¡ +¢ <br />

∆ ‡ exp<br />

µ− ∆+‡<br />

exp<br />

<br />

• for unimolecular reactions:<br />

( )= µ<br />

<br />

exp<br />

• for bimolecular reactions:<br />

( )= µ<br />

<br />

+ exp<br />

− ∆+‡<br />

µ− ∆+‡<br />

<br />

− ∆+‡<br />

<br />

<br />

<br />

µ− ∆+‡<br />

<br />

exp<br />

<br />

exp<br />

<br />

<br />

+‡<br />

∆<br />

exp<br />

µ−<br />

<br />

<br />

µ− ∆+‡<br />

<br />

µ− ∆+‡<br />

<br />

<br />

<br />

(7.93)<br />

(7.94)<br />

(7.95)<br />

(7.96)<br />

Note: The conversion between values for ∆ +‡ and ∆ ª‡ and, likewise, between<br />

values for ∆ +‡ and ∆ ª‡ for a given reaction in the liquid and gas phase requires<br />

some caution. 46<br />

46 See,e.g.,D.M.Golden,J.Chem.Ed.48, 235 (1971).


7.3 Thermodynamic interpretation of transition state theory 164<br />

7.3.2 Applications of thermodynamic transition state theory<br />

a) Comparison of ∆S ‡ -values for different types of transition states<br />

I<br />

Bimolecular reactions:<br />

A+BC→ ABC ‡ (7.97)<br />

y<br />

∆ ‡ 0 (7.98)<br />

or even<br />

∆ ‡ ¿ 0 (7.99)<br />

Furthermore, for chemically otherwise similar molecules<br />

(nonlinear molecule) (linear molecule) À (cyclic molecule) (7.100)<br />

Thus,<br />

¯<br />

¯∆ ‡ (nonlinear molecule)¯¯ ¯¯∆ ‡ (linear molecule)¯¯ ¿ ¯¯∆ ‡ (cyclic molecule)¯¯<br />

(7.101)<br />

For example:<br />

⎛<br />

A ··· B<br />

∆ ‡ ⎜<br />

⎝<br />

¯<br />

y<br />

...<br />

⎞ ‡ ⎛<br />

‡ ⎞ ⎛<br />

⎟ A ··· B<br />

<br />

¯<br />

C<br />

⎠¯¯¯¯¯¯¯ ¯∆‡ ⎝ A ··· B ··· C ⎠¯<br />

⎞ ‡ ¯ ¿<br />

∆ ‡ ⎜ ⎟<br />

⎝<br />

. .<br />

¯ C ··· D<br />

⎠¯¯¯¯¯¯¯<br />

(7.102)<br />

cyclic ¿ linear nonlinear (7.103)<br />

I<br />

Unimolecular reactions:<br />

ABC → ABC ‡ (7.104)<br />

• For isomerization reactions with a tight transition state:<br />

∆ ‡ ⎛<br />

⎝<br />

• For bond fission reactions with a loose transition state:<br />

‡<br />

B<br />

⎞<br />

Á Â ⎠ . 0 (7.105)<br />

A – C<br />

∆ ‡ ¡ AB ···C ‡¢ 0 (7.106)<br />

y<br />

loose tight (7.107)


7.3 Thermodynamic interpretation of transition state theory 165<br />

b) Thermochemical estimation of A-factors*<br />

Equations 7.71 and 7.86 allow us to estimate the -factors of chemical reaction rate<br />

constants by the following procedure (Benson1976):<br />

(1) We setup a model for the TS, with a certain structure and vibrational freuqencies.<br />

(2) We take a reference molecule that is similar to the TS and use the ref of the<br />

reference molecule for a zero-order estimation of ∆ ‡ according to<br />

∆ ‡ = ref −<br />

X<br />

( reactants ) (7.108)<br />

reactants<br />

(3) We apply corrections for the differences between ref and ‡ basedonthestatistical<br />

thermodynamics result of<br />

= ln + ln <br />

(7.109)<br />

<br />

y<br />

µ <br />

∆ cor = ‡ − ref = ln ‡ <br />

‡<br />

+ <br />

ref ln (7.110)<br />

ref<br />

Usually, the correction term accounts for differences of the entropies for translation,<br />

rotation, internal rotations, vibrations, symmetry, electron spin, plus sometimes<br />

other terms (e.g., optical isomers).<br />

I Example 7.1: Estimation of the -factor for the reaction O+CH 4 →<br />

(O ···H ···CH 3 ) ‡ → OH + CH 3 .<br />

∆ ‡ = (O ···H ···CH 3 ) ‡ − (O) − (CH 4 ) (7.111)<br />

ref = (CH 3 F) (7.112)<br />

∆ cor thus depends on a comparison of CH 3 F and OHCH ‡ 3 (see Table ). ¤<br />

I Table 7.4: Estimation of the -factor for the reaction O+CH 4 →<br />

(O ···H ···CH 3 ) ‡ → OH + CH 3 .<br />

Contribution ∆ ‡ (Jmol −1 K −1 )<br />

ª (CH 3F) − ª (O) − ª (CH 4)<br />

−1243<br />

Spin: + ln 3 + 91<br />

external rotation: 1 3 ln ( 1 2 3 ) ‡<br />

( 1 2 3 ) ref + 79<br />

translation − 04<br />

CF =1000cm −1 → − 04<br />

OH =2000cm −1 + 00<br />

2 OHC ()<br />

³<br />

=600cm −1 ´<br />

+ 42<br />

∆ ‡ = ª <br />

O ···H ···CH ‡ 3 − ª (O) − ª (CH 4) −1039<br />

Correction ∆ ‡ → ∆ ‡ : + ln ( 0 ) y ∆ ‡ = −198 Jmol −1 K −1 .


7.3 Thermodynamic interpretation of transition state theory 166<br />

Estimated -value:<br />

= 2 µ <br />

∆<br />

‡<br />

exp (7.113)<br />

<br />

µ −197Jmol −1<br />

=739 × 625 × 10 12 K −1 <br />

× exp<br />

831 J mol −1 K −1 cm 3 mol −1 s −1 (7.114)<br />

=43 × 10 12 cm 3 mol −1 s −1 (7.115)<br />

Experimental -value:<br />

expt =19 × 10 12 cm 3 mol −1 s −1 (7.116)<br />

The difference is probably due to a poor estimate of the TS structure <strong>—</strong> one can do<br />

much better, I just didn’t try hard enough here. 47<br />

7.3.3 Kinetic isotope effects<br />

I<br />

Figure 7.5: The kinetic isotope effect due to the difference in the zero-point energies<br />

of the reactants and the respective TS structures.<br />

47 For H abstraction reactions from a series of hydrocarbons by 3 CH 2 , the agreement that was achieved<br />

was better than ±30 %.


7.3 Thermodynamic interpretation of transition state theory 167<br />

Kinetic isotope effects occur in particular in the case of H/D atom transfer reactions.<br />

The reason is that the threshold energy ∆ ‡ 0 depends on the vibrational zero-point<br />

energies.<br />

We consider the reactions<br />

where<br />

A+H<strong>—</strong>R → A<strong>—</strong>H<strong>—</strong>R ‡ → products (7.117)<br />

A+D<strong>—</strong>R → A<strong>—</strong>D<strong>—</strong>R ‡ → products (7.118)<br />

∆ ‡ 0 (H<strong>—</strong>R) = ∆ ‡ + ‡ (A<strong>—</strong>H<strong>—</strong>R) − (H<strong>—</strong>R) (7.119)<br />

∆ 0(D<strong>—</strong>R) ‡ = ∆ ‡ + ‡ (A<strong>—</strong>D<strong>—</strong>R) − (D<strong>—</strong>R) (7.120)<br />

with<br />

y<br />

= <br />

1<br />

2<br />

X<br />

(7.121)<br />

<br />

∆ ‡ 0(DH) = ∆‡ 0(D<strong>—</strong>R) − ∆ 0(H<strong>—</strong>R) ‡ (7.122)<br />

=+ ‡ (A<strong>—</strong>D<strong>—</strong>R) − ‡ (A<strong>—</strong>H<strong>—</strong>R) − ( (D<strong>—</strong>R) − (H<strong>—</strong>R))<br />

(7.123)<br />

y<br />

à !<br />

(D<strong>—</strong>R)<br />

(H<strong>—</strong>R) =exp − ∆‡ 0(D/H)<br />

<br />

(7.124)<br />

I<br />

Example 7.2: R-H/R-D substitution. The vibrational frequencies depend on the force<br />

constants and reduced masses :<br />

= 1 µ 12 <br />

(7.125)<br />

2 <br />

For H/D substitution:<br />

(HR)<br />

(DR) ≈ 1 (7.126)<br />

2<br />

y<br />

µ 12<br />

(HR) (HR)<br />

(DR) = ≈ √ 2 (7.127)<br />

(DR)<br />

With (HR) = 3000 cm −1 and (DR) = 2100 cm −1 , we find ∆ =<br />

1<br />

2 (3000 − 2100) cm−1 =450cm −1 , by which the DR reactant is stabilized compared<br />

to the HR reactant. For =300K:<br />

µ<br />

(DR)<br />

(HR) =exp − 450 <br />

≈ −225 ≈ 01 (7.128)<br />

200<br />

For more realistic estimates, one has to take into account the change of all vibrational<br />

frequencies in the reacting molecules and the TS.<br />

¤


7.3 Thermodynamic interpretation of transition state theory 168<br />

7.3.4 Gibbs free enthalpy correlations<br />

I<br />

Figure 7.6: Linear free enthalpy correlations.<br />

Application of the TST expression<br />

= µ <br />

<br />

exp − ∆‡<br />

(7.129)<br />

<br />

to a series of related reactions (1) and (2) often gives a relation of the type 48<br />

∆ ‡ 1 − ∆ ‡ 2 = ¡ ¢<br />

∆ ª 1 − ∆ ª 2<br />

(7.130)<br />

y<br />

ln 1<br />

= ln 1<br />

(7.131)<br />

2 2<br />

7.3.5 Pressure dependence of k<br />

Liquid phase reactions often show a pressure dependence of the rate constant. Using<br />

the relation<br />

µ <br />

= (7.132)<br />

<br />

<br />

we can predict this pressure dependence. Starting with<br />

= µ <br />

<br />

exp − ∆‡<br />

(7.133)<br />

<br />

we find<br />

µ ln <br />

= − ∆ ‡<br />

(7.134)<br />

<br />

<br />

48 A well-known example is the Hammett correlation in organic chemistry.


7.3 Thermodynamic interpretation of transition state theory 169<br />

I<br />

Activation volume:<br />

∆ ‡ (7.135)<br />

I Example 7.3: Comparison of the pressure dependencies of the rate constants for 1<br />

and 2 reactions:<br />

(CH 3<br />

) 3<br />

CCl + H 2 O → (CH 3<br />

) 3<br />

C + +Cl − +H 2 O → ∆ ‡ 0<br />

y ↑ y ↓ 1<br />

HO − +CH 3 I<br />

⎡<br />

⎤<br />

|<br />

→ ⎣ HO ···C ···I ⎦<br />

ÁÂ<br />

−<br />

→ ∆ ‡ 0<br />

y ↑ y ↑ 2<br />

¤


7.4 Transition state spectroscopy 170<br />

7.4 Transition state spectroscopy<br />

Until a few years ago, the “transition state” was a rather “elusive and strange” species.<br />

This situation has completely changed (a) because of the rapidly improving quantum<br />

chemical methods for computing TS properties and (b) because of transition state<br />

spectroscopy, in the frequency domain (J. Polanyi, Kinsey), and in the time domain<br />

(Zewail).<br />

I<br />

Figure 7.7: Transition state spectroscopy: The reaction K + NaCl.<br />

I<br />

Figure 7.8: Transition state spectroscopy: Raman emission of dissociating molecules.


7.4 Transition state spectroscopy 171<br />

I Figure 7.9: Transition state spectroscopy: The reaction F + H 2 .<br />

I<br />

Figure 7.10: Transition state spectroscopy: NaI Photodissociation.


7.4 Transition state spectroscopy 172<br />

I<br />

Figure 7.11: Transition state spectroscopy: NaI Photodissociation.<br />

I<br />

Figure 7.12: Transition state spectroscopy: Reaction a in van-der-Waals Complex.


7.4 Transition state spectroscopy 173<br />

I<br />

Figure 7.13: Transition state spectroscopy: Stimulated Emission Pumping.


7.4 Transition state spectroscopy 174<br />

7.5 References<br />

Benson 1976 S. W. Benson, Thermochemical <strong>Kinetics</strong>, Methods for the Estimation<br />

of Thermochemical Data and Rate Parameters, 2nd Ed., Wiley, New York, 1976.<br />

Evans 1935 M. G. Evans, M. Polanyi, Trans. Faraday Soc. 31, 875 (1935).<br />

Eyring 1935 H. Eyring, J. Chem. Phys. 3, 107 (1935).<br />

Herschbach 1987 D. R. Herschbach, Molekulardynamik einfacher chemischer Reaktionen,Angew.Chem.99,<br />

1251 (1987).<br />

Lee 1987 Y. T. Lee, Molekularstrahluntersuchungen chemischer Elementarprozesse,<br />

Angew. Chem. 99, 967 (1987).<br />

Polanyi 1987 J. C. Polanyi, Konzepte der Reaktionsdynamik, Angew. Chem. 99, 981<br />

(1987).<br />

Wigner 1938 E. Wigner, Trans. Faraday Soc. 34, 29 (1938).


7.5 References 175<br />

8. Unimolecular reactions<br />

Unimolecular reactions are reactions in which only one species experiences chemical<br />

change.<br />

I<br />

Figure 8.1: Examples of unimolecular reactions.<br />

8.1 Experimental observations<br />

(1) There are many reactions governed by a first-order rate law,<br />

[A]<br />

= − ∞ [A] , (8.1)<br />

<br />

for example isomerization and dissociation reactions like<br />

Cyclopropane → Propene (8.2)<br />

-Butene → -Butene (8.3)<br />

C 2 H 6 → 2CH 3 (8.4)<br />

Cycloheptatriene → Toluene (8.5)<br />

(2) In these reactions, no other species than the reactants experience any chemical<br />

change.<br />

(3) These reactions have the same rates in the gas phase and in solution,<br />

gas phase ≈ solution phase , (8.6)<br />

i.e., the environment has no effect on the reaction rates!


8.1 Experimental observations 176<br />

(4) There are many similar other reactions, however, for which the rate law is secondorder,<br />

i.e.,<br />

[A]<br />

= − 0 [A] [M] (8.7)<br />

<br />

Examples are<br />

Br 2 → Br + Br (8.8)<br />

H 2 O 2 → OH + OH (8.9)<br />

(5) More detailed measurements show that at a given temperature the order of the<br />

reaction may depend on pressure (see Fig. 8.2), for instance for the reaction<br />

one has found that<br />

CH 3 NC → CH 3 CN (8.10)<br />

a) for → 0:<br />

[CH 3 NC]<br />

= − 0 [CH 3 NC] [M] (8.11)<br />

<br />

This is called the “low pressure range” for the reaction.<br />

b) for →∞:<br />

[CH 3 NC]<br />

= − ∞ [CH 3 NC] (8.12)<br />

<br />

This is called the “high pressure range” for the reaction”.<br />

(6) The transition from the low to the high pressure range depends<br />

a) onthesizeofthemolecule(seeFig.8.3),<br />

b) for a given molecule, it depends on temperature.<br />

(7) The activation energy in the low pressure regime is much smaller than in the high<br />

pressure regime:<br />

= 2 ln <br />

(8.13)<br />

<br />

where<br />

0 ∞ (8.14)<br />

and<br />

∞ ≈ 0 (8.15)<br />

(8) The preexponential factors in the low pressure regime are much higher than the<br />

collision frequencies, i.e.,<br />

0 (8.16)<br />

(9) For all these reactions, the molecules must somehow be “energized”. There are<br />

other reactions, in which the molecules are energized chemically (“chemically<br />

activated reactions”) or by photons. At the microscopic, i.e., molecular level,<br />

these reactions should have related mechanisms.


8.1 Experimental observations 177<br />

I<br />

Questions:<br />

(1) What are the differences of the reactions under (1), (4) and (5)?<br />

(2) How can we rationalize their reaction mechanisms?<br />

I<br />

Figure 8.2: Schematic fall-off curve of the unimolecular rate constant.<br />

I<br />

Figure 8.3: Experimentally observed fall-off curves of unimolecular reactions.


8.2 Lindemann mechanism 178<br />

8.2 Lindemann mechanism<br />

First attempts to explain the pressure dependence of unimolecular reaction rate constants<br />

due to radiation and the dissociation dynamics due to centrifugal forces (rotational<br />

excitation) proved to be wrong. In addition, the historic examples for unimolecular reactions,<br />

for instance the N 2 O 5 decomposition, were later proven to have more complex<br />

reaction mechanisms. The first “realistic” model that explained the pressure dependence<br />

of was that proposed by Lindemann (1922).<br />

I<br />

Figure 8.4: Lindemann mechanism.<br />

I<br />

Figure 8.5: Lindemann mechanism: High pressure regime.


8.2 Lindemann mechanism 179<br />

I<br />

Figure 8.6: Lindemann mechanism: Low pressure regime.<br />

I<br />

Figure 8.7: Fall-off curve: The proper way to plot log is vs. log [M]. ab<br />

a<br />

Since 0 ∝ [M], wecanalsoplotlog vs. log 0 . This gives the so-called doubly reduced fall-off<br />

curves.<br />

b In reality, the fall-off regime extends over several orders of magnitude in [M].


8.2 Lindemann mechanism 180<br />

I Half-pressure: We have found the following results:<br />

=<br />

2 1 [M]<br />

−1 [M] + 2<br />

(8.17)<br />

Thus, if −1 [M] = 2 ,wehave<br />

1<br />

∞ = 2<br />

−1<br />

(8.18)<br />

0 = 1 [M] (8.19)<br />

=<br />

2 1 [M]<br />

= 2 1 [M]<br />

−1 [M] + 2 2 −1 [M] = 1 2 1<br />

= 1 2 −1 2 ∞ (8.20)<br />

Half “pressure”:<br />

[M] 12<br />

= 2<br />

−1<br />

(8.21)<br />

I Lifetimes of energized molecules: It is instructive to look at the reciprocal values:<br />

1<br />

−1 [M] = 1 2<br />

(8.22)<br />

1<br />

• is the time between two collisions. This is, in fact, the lifetime of the<br />

−1 [M]<br />

energized molecules with respect to collisional deactivation.<br />

• 1 is the lifetime of the excited molecules with respect to reaction.<br />

2<br />

Thus, at the half pressure, we have<br />

Collisional Deactivation = Unimolecular Decay (8.23)<br />

I Determination of k ∞ :<br />

y<br />

=<br />

2 1 [M]<br />

−1 [M] + 2<br />

(8.24)<br />

1<br />

= −1<br />

+ 1 1<br />

1 2 1 [M]<br />

(8.25)<br />

= 1 + 1 1<br />

∞ 1 [M]<br />

(8.26)<br />

= 1 + 1 ∞ 0<br />

(8.27)<br />

y Plot of −1 vs. [M] −1 should give a straight line with slope 1 −1 and intercept<br />

∞ −1 .


8.2 Lindemann mechanism 181<br />

I<br />

Deficiencies of the Lindemann model:<br />

I Figure 8.8:<br />

• Activation energies:<br />

In particular:<br />

0 ∞ (8.28)<br />

∞ ≈ 0 (8.29)<br />

0 ≈ 0 − ( − 15) (8.30)


8.3 Generalized Lindemann-Hinshelwood mechanism 182<br />

8.3 Generalized Lindemann-Hinshelwood mechanism<br />

8.3.1 Master equation<br />

I<br />

Figure 8.9: Strong collision model.<br />

I<br />

Figure 8.10: Generalized Lindemann-Hinshelwood weak collision model.


8.3 Generalized Lindemann-Hinshelwood mechanism 183<br />

I<br />

Figure 8.11: Generalized Lindemann-Hinshelwood weak collision model.<br />

I<br />

Figure 8.12: Generalized Lindemann-Hinshelwood model: Rate constants in the high<br />

and low pressure (strong and weak collision) limits.


8.3 Generalized Lindemann-Hinshelwood mechanism 184<br />

8.3.2 Equilibrium state populations<br />

As seen from the master equation analysis, the unimolecular reaction rate constant<br />

in the high and in the low pressure regimes depends on the equilibrium (Boltzmann)<br />

state distributions (for discrete states) or () (continuous state distributions). In<br />

a polyatomic molecule, due to the large number of vibrational degrees of freedom, the<br />

number of vibrational states increases very rapidly with increasing vibrational excitation<br />

energy. The quanta of vibrational excitation can be distributed over the oscillators. In<br />

order to calculate the number of combinations, we start with a single oscillator and<br />

expand our scope first to two oscillators and then to =3 − 6 oscillators.<br />

a) Equilibrium state population for a single oscillator:<br />

I<br />

Boltzmann distribution:<br />

with the energy quanta<br />

the degeneracy factor<br />

and the partition function<br />

= − <br />

<br />

(8.31)<br />

= × ; =0 1 2 (8.32)<br />

=<br />

(8.33)<br />

1<br />

1 − − <br />

(8.34)<br />

I Limiting case for high T (transition to classical mechanics): ¿ y<br />

classical<br />

<br />

= <br />

<br />

(8.35)<br />

since<br />

− ≈ 1 − for → 0 (8.36)<br />

Note that this classical description of vibrations is not so bad for unimolecular reactions,<br />

where we are interested in the kinetic behavior at high temperatures.<br />

b) Equilibrium state population for s oscillators:<br />

y<br />

() → () (8.37)<br />

() = () − <br />

(8.38)<br />

<br />

with () being the density of vibrational states at the energy and<br />

=<br />

Y<br />

=1<br />

1<br />

1 − − <br />

(8.39)<br />

and<br />

=3 − 6 (resp. =3 − 5) (8.40)


8.3 Generalized Lindemann-Hinshelwood mechanism 185<br />

8.3.3 The density of vibrational states ρ (E)<br />

Thedensityofvibrationalstatesisdefined as<br />

() =<br />

number of states in energy interval<br />

energy interval<br />

(8.41)<br />

i.e.,<br />

() =<br />

()<br />

<br />

(8.42)<br />

In a polyatomic molecule, as we shall see in the following, () increases with very<br />

rapidly owing to the large number of overtone and combination states, and reaches truly<br />

astronomical values:<br />

a) s =1:<br />

() = 1<br />

<br />

(8.43)<br />

b) s =2:<br />

We consider the ways, in which we can distribute two identical energy quanta between<br />

the two oscillators:<br />

Osc. 1 Osc. 2 Osc. 1 Osc. 2 Osc. 1 Osc. 2 Osc. 1 Osc. 2 <br />

0 - - 1<br />

| - - | 2<br />

2 || - | | - || 3<br />

3 ||| - || | | || - ||| 4<br />

4 5<br />

5 6<br />

.<br />

.<br />

.<br />

.<br />

(8.44)<br />

c) s oscillators:<br />

• For simplicity, we consider a molecule with identical harmonic oscillators.<br />

• This molecule shall be excited with quanta, eachofsize, i.e., the molecule<br />

has the excitation energy = <br />

I<br />

Question: How can we distribute these quanta over the oscillators?


8.3 Generalized Lindemann-Hinshelwood mechanism 186<br />

I<br />

Answer: We are asking for the number of distinguishable possibilities to distribute <br />

quanta between oscillators, (). This is identical to the problem of distributing <br />

balls between boxes, for example for =11and =8:<br />

••• • • • •• • •• (8.45)<br />

That number is identical to the number of distinguishable permutations of balls ( • )<br />

and − 1 walls ( | ),<br />

( + − 1)!<br />

() = (8.46)<br />

!( − 1)!<br />

where ( + − 1)! is the total number of permutations, which is corrected by dividing<br />

through !( − 1)! to account for the number of indistinguishable permutations among<br />

the balls (!) andwalls(( − 1)!) among each other (which give indistinguishable results).<br />

I Answer: We make the following approximation for À : 49<br />

( + − 1)!<br />

!( − 1)!<br />

≈<br />

−1<br />

( − 1)!<br />

(8.49)<br />

I<br />

Question: How can we compute ()?<br />

I<br />

Answer:<br />

() =<br />

()<br />

<br />

(8.50)<br />

(1) We consider the energy interval<br />

∆ = (8.51)<br />

at the excitation energy<br />

= × (8.52)<br />

In this energy interval, we have the number of states<br />

∆() = () (8.53)<br />

y<br />

() =<br />

()<br />

<br />

∆ ()<br />

≈<br />

∆ = 1 ( + − 1)!<br />

!( − 1)!<br />

(8.54)<br />

49 The approximation<br />

( + − 1)!<br />

≈ −1 (8.47)<br />

!<br />

can be derived using Stirling’s formula (ln ! = ln − ) and the power series expansion of<br />

for || ¿1 (see homework assignment).<br />

ln (1 − ) ≈ (8.48)


8.3 Generalized Lindemann-Hinshelwood mechanism 187<br />

(2) Using the approximation<br />

( + − 1)!<br />

!( − 1)!<br />

≈<br />

−1<br />

( − 1)!<br />

(Eq. 8.49) for À gives<br />

() = 1 −1<br />

( − 1)!<br />

(8.55)<br />

(3) Inserting = into (), weobtain<br />

() = 1 () −1<br />

( − 1)!<br />

(8.56)<br />

(4) Result:<br />

() =<br />

−1<br />

( − 1)! () (8.57)<br />

d) Notes:*<br />

(1) The above derivation may seem artificial because in a real molecule not all oscillators<br />

are identical. However, the derivation can be easily generalized.<br />

(2) For instance, we can start from the number of states of the first oscillator 1 ( 1 )<br />

and convolute this with the density of states 2 () of the second oscillator<br />

2 ( 2 )= 2 ( − 1 )= 1<br />

(8.58)<br />

<br />

and compute the total (combined) density of states for two oscillators by convolution<br />

according to<br />

() =<br />

Z <br />

etc. up to oscillators. The result is<br />

0<br />

1 ( 1 ) 2 ( − 1 ) 1 (8.59)<br />

() =<br />

−1<br />

( − 1)! Q <br />

=1 <br />

(8.60)<br />

(3) We can also use inverse Laplace transforms: Since is the Laplace transform<br />

of () according to<br />

=<br />

Z ∞<br />

0<br />

() − = L [ ()] (8.61)<br />

we can determine () from (which we know) by inverse Laplace transformation.<br />

(4) With corrections for zero-point energy:<br />

() =<br />

( + ) −1<br />

( − 1)! Q <br />

=1 <br />

(8.62)


8.3 Generalized Lindemann-Hinshelwood mechanism 188<br />

(5) With (empirical) Whitten-Rabinovitch corrections:<br />

() = ( + () ) −1<br />

( − 1)! Q <br />

=1 <br />

(8.63)<br />

where () is a correction factor that is of the order of 1 except at very low<br />

energies.<br />

(6) Exact values of () for harmonic oscillators can be obtained by direct state<br />

counting algorithms.<br />

(7) Corrections for anharmonicity can be applied using different means.<br />

8.3.4 Unimolecular reaction rate constant in the low pressure regime<br />

From the master equation analysis above, we obtained the thermal unimolecular reaction<br />

rate constant in the low pressure regime as<br />

0 = X X<br />

−1() [M] <br />

(8.64)<br />

<br />

<br />

<br />

The reduction of the excited state populations compared to the equilibrium (Boltzmann)<br />

distribution is shown in Fig. 8.13. Two cases can be distinguished:<br />

(1) With the strong collision assumption (h∆ iÀ), we can apply the so-called<br />

equilibrium theories which assume that = for ≤ 0 . Thus, the state<br />

population below 0 remains as in the high pressure regime, whereas it is reduced<br />

above 0 .Then,<br />

y<br />

0 = X <br />

X<br />

<br />

−1() [M] <br />

<br />

= X <br />

Z ∞<br />

0 = [M]<br />

X<br />

−1() [M] = [M] X<br />

<br />

<br />

(8.65)<br />

0<br />

() (8.66)<br />

Using the expressions for () () and in the classical limit ( ¿ )<br />

derived above and doing the integration by parts, we obain the Hinshelwood<br />

equation for 0 :<br />

µ −1 0 1<br />

0 = [M]<br />

( − 1)! − 0 <br />

(8.67)<br />

This equation is usually used to describe experimental data with as a fit parameter<br />

(usually about half the real ).<br />

Since ∝ 05 and = 2 ln , the low pressure activation energy<br />

becomes<br />

0 = 0 − ( − 15) (8.68)<br />

Thus, may be significantly smaller than 0 . This can be understood as<br />

illustrated in Fig. 8.14.


8.3 Generalized Lindemann-Hinshelwood mechanism 189<br />

(2) With the weak collision assumption (h∆ i ≤ ) made by the so-called nonequilibrium<br />

theories, the bottleneck in the state population falls below the equilibrium<br />

distribution already below 0 (see Fig. 8.14). Using the weak collision<br />

assumption, Troe derived the expression<br />

0 = [M] ( 0 )<br />

<br />

− 0 <br />

(8.69)<br />

where is the so-called collision efficiency ( ≤ 1) whichcanbeexpressedin<br />

terms of the average energy transferred per collision h∆ i. becomes even<br />

smaller than the Hinshelwood above. A simple estimate of the weak collision<br />

effect based on Fig. 8.14 may lead to<br />

0 ≈ 0 − ( − 05) (8.70)<br />

I<br />

Figure 8.13: State populations.


8.3 Generalized Lindemann-Hinshelwood mechanism 190<br />

I<br />

Figure 8.14: Activation energies for unimolecular reactions in the high pressure, the<br />

low pressure strong collision, and the low pressure weak collision limits.<br />

8.3.5 Unimolecular reaction rate constant in the high pressure regime<br />

From the master equation analysis above, the thermal unimolecular reaction rate constant<br />

in the high pressure regime is (see Fig. 8.12)<br />

with the Boltzmann distribution<br />

asshowninFig.8.13.<br />

∞ =<br />

Z ∞<br />

0<br />

() () (8.71)<br />

() = () − <br />

<br />

(8.72)<br />

Averaging over the Boltzmann distribution gives the TST expression<br />

( )= <br />

<br />

‡<br />

− 0 <br />

(8.73)


8.4 The specific unimolecular reaction rate constants () 191<br />

8.4 The specific unimolecular reaction rate constants k (E)<br />

We may intuitively expect that the specific uinimolecular reaction rate constants ()<br />

depend on the energy content (i.e., the excitation energy ) of the energetically activated<br />

molecules molecules A ∗ . More precisely, we will also have to take into account<br />

the dependence of the specific rate constants on other conserved degrees of freedom,<br />

in particular total angular momentum (⇒ ()), and perhaps other quantities<br />

(symmetry, components of if -mixing is weak, (⇒ (; ; ))).<br />

8.4.1 Rice-Ramsperger-Kassel (RRK) theory<br />

I RRK model: In order to derive an expression for the specific rateconstants (),<br />

werealizethatitisnotenoughthatthemoleculesareenergizedtoabove 0 . Indeed,<br />

for the reaction to take place, the energy also has to be in the right oscillator, namely<br />

in the RC. In particular, for the reaction to take place, of the total excitation energy ,<br />

the amount † has to be concentrated in the RC such that † ≥ 0 . Thus, we can<br />

write the reaction scheme as<br />

∗ †<br />

→ † → (8.74)<br />

where † denotes those excited molecules that have enough energy in the RC to overcome<br />

0 . We assume that once this critical configuration ( † ) is reached, the molecules<br />

will immediately cross the TS and go to the product side.<br />

I Probabilities of A † and A ∗ : The probability of † compared to ∗ can be derived<br />

using statistical arguments by considering the ratio of the density of states. We consider<br />

amoleculewith<br />

• oscillators,<br />

• = quanta which have to be distributed over the oscillators,<br />

• a threshold energy of 0 such that a minimum of = 0 quanta have to be<br />

concentrated in the RC and only − quanta can be distributed freely,<br />

• À and − À .<br />

The probability of † relative to ∗ is given by<br />

¡ †¢<br />

( ∗ ) = ¡ †¢<br />

( ∗ )<br />

(8.75)<br />

where<br />

( ∗ ) ∝ ∗ () = 1 ( + − 1)!<br />

(8.76)<br />

!( − 1)!<br />

and, since in the critical configuration only − quanta can be distributed freely,<br />

¡ †¢ ∝ † () = 1 ( − + − 1)!<br />

( − )! ( − 1)!<br />

(8.77)


8.4 The specific unimolecular reaction rate constants () 192<br />

y<br />

¡ †¢<br />

( ∗ )<br />

( − + − 1)! !( − 1)!<br />

= ×<br />

( − )! ( − 1)! ( + − 1)!<br />

( − + − 1)! !<br />

= ×<br />

( − )! ( + − 1)!<br />

(8.78)<br />

(8.79)<br />

I Specific rate constant: The critical configuration † is reached with the rate coefficient<br />

(≡ frequency factor) † (see the reaction scheme above). Since the energy is now<br />

in place (in the RC), † will cross the TS to products.<br />

The specific rateconstant () is therefore simply<br />

y<br />

() = † ×<br />

() = † × ¡ †¢<br />

( ∗ )<br />

( − + − 1)!<br />

( − )!<br />

×<br />

!<br />

( + − 1)!<br />

(8.80)<br />

(8.81)<br />

I Kassel formula: With the approximations according to Eq. 8.49),<br />

( + − 1)!<br />

!<br />

≈ −1 for À (8.82)<br />

and<br />

we obtain<br />

( − + − 1)!<br />

( − )!<br />

¡ †¢<br />

( ∗ )<br />

≈ ( − ) −1 for − À , (8.83)<br />

=<br />

( − )−1<br />

−1 =<br />

µ −1 − <br />

(8.84)<br />

y<br />

¡ †¢ µ −1 −<br />

( ∗ ) = 0<br />

(8.85)<br />

<br />

so that the specific rateconstantbecomes<br />

<br />

This expression is known as the Kassel formula for ().<br />

µ −1 −<br />

() = † 0<br />

(8.86)<br />

<br />

I<br />

Notes on the Kassel formula:<br />

• Owing to the approximations À and − À , the Kassel formula is valid<br />

only at À 0 .


8.4 The specific unimolecular reaction rate constants () 193<br />

• The Kassel formula describes experimental data qualitatively well. However, instead<br />

of the full value of =3 − 6, one usually has to use an effective number<br />

eff . As suggested by Troe, eff can be estimated from the specific heatofthe<br />

reactant molecules via<br />

=<br />

µ <br />

<br />

<br />

<br />

= <br />

<br />

<br />

µ<br />

2 ln <br />

<br />

using the known expression for ,<br />

Y 1<br />

=<br />

1 − exp (− )<br />

=1<br />

<br />

≈ eff (8.87)<br />

(8.88)<br />

• Often, eff is equal to about one half of the number of oscillators,<br />

eff ≈<br />

(3 − 6)<br />

2<br />

(8.89)<br />

I<br />

Figure 8.15: Specific rate constants: RRK formula.<br />

8.4.2 Rice-Ramsperger-Kassel-Marcus (RRKM) theory<br />

RRK theory fails in the following points:<br />

• Itcanbeusedonlywith as an effective parameter ( eff ).<br />

• It leads to the wrong results close to threshold, where → 0 (see the conditions<br />

above).<br />

• It is difficult to derive for oscillators with different frequencies.<br />

• RRK theory is essentially a classical theory. It does not know about zero-point<br />

energy.


8.4 The specific unimolecular reaction rate constants () 194<br />

I RRKM expression: These problems are essentially solved by RRKM theory which<br />

gives the expression<br />

() = ‡ ( − 0 )<br />

()<br />

(8.90)<br />

where<br />

• ‡ ( − 0 ) is the number of open channels (essentially the number of states at<br />

the transition state configuration exluding the RC), counted from the zero-point<br />

level of the TS, at the energy − 0 ,<br />

• () is the density of states of the reacting molecules at the energy .<br />

These quantities are the two critical quantities in microcanonical transition state theory.<br />

I<br />

Figure 8.16: RRKM expression for ().


8.4 The specific unimolecular reaction rate constants () 195<br />

I<br />

Figure 8.17: RRKM expression for () with angular momentum conservation.<br />

I<br />

Figure 8.18: The sum and the density of states.


8.4 The specific unimolecular reaction rate constants () 196<br />

I RRKM expression with angular momentum conservation: Eq. 8.90 can be generalized<br />

to include other conserved quantities, e.g., total angular momentum,<br />

() = ‡ ( − 0 )<br />

()<br />

(8.91)<br />

I Relation between RRKM and RRK theory: To elucidate the relation between Eqs.<br />

8.86 and 8.90, we look at () and ( − 0 ) more closely.<br />

(1) Density of vibrational states ():<br />

() =<br />

number of states in energy interval<br />

energy interval<br />

=<br />

()<br />

<br />

(8.92)<br />

a) RRK result for the number of indistinguishable permutations of quanta<br />

over oscillators; = À :<br />

y<br />

() = 1 ( + − 1)!<br />

!( − 1)!<br />

() =<br />

≈ 1 −1<br />

( − 1)! = 1 () −1<br />

( − 1)!<br />

(8.93)<br />

−1<br />

( − 1)! () (8.94)<br />

b) It is relatively easy to show that for oscillators with different frequencies, the<br />

expression for () becomes<br />

() =<br />

−1<br />

( − 1)! Q <br />

=1 <br />

(8.95)<br />

c) With corrections for zero-point energy:<br />

() =<br />

( + ) −1<br />

( − 1)! Q <br />

=1 <br />

(8.96)<br />

d) With (empirical) Whitten-Rabinovitch correction:<br />

() = ( + () ) −1<br />

( − 1)! Q <br />

=1 <br />

(8.97)<br />

where () is a correction factor that is of the order of 1 except at very<br />

low energies.<br />

e) Exact values of () for harmonic oscillators can be obtained by direct<br />

state counting algorithms. Corrections for anharmonicity can be applied<br />

using different means.


8.4 The specific unimolecular reaction rate constants () 197<br />

(2) Number of states () and ‡ ( − 0 ): Since () = () , the<br />

number of states from =0to can be obtained by integration,<br />

a) RRK expression:<br />

y<br />

() =<br />

Z <br />

0<br />

() =<br />

() =<br />

Z <br />

0<br />

Z <br />

0<br />

() (8.98)<br />

−1<br />

( − 1)! () =<br />

() =<br />

Z<br />

1 1<br />

( − 1)! () −1 <br />

0<br />

(8.99)<br />

<br />

!() (8.100)<br />

b) Allowing for different frequencies, we modify this expression to<br />

() =<br />

<br />

! Q <br />

=1 <br />

(8.101)<br />

c) As above, a correction for zero-point energy gives<br />

() = ( + ) <br />

! Q <br />

=1 <br />

(8.102)<br />

d) With the Whitten-Rabinovitch correction, this becomes<br />

() = ( + () ) <br />

! Q <br />

=1 <br />

(8.103)<br />

e) Correction for → 0: At =0,wemusthaveatleastonestate,i.e.,<br />

‡ () =1+ ( + () ) <br />

! Q <br />

=1 <br />

− ( (0) ) <br />

! Q <br />

=1 <br />

(8.104)<br />

(3) Application to unimolecular reactions:<br />

a) At the TS, we have only − 1 oscillators since the RC has to be excluded.<br />

Furthermore, the states range from the zero-point energy 0 of the TS to<br />

, while we count from the zero-point of the reactants. y<br />

‡ ( − 0 )=1+<br />

³<br />

− 0 + ‡ ( − 0 ) ‡ ´−1<br />

<br />

−<br />

³<br />

‡ (0) ‡ ´−1<br />

<br />

( − 1)! Q −1<br />

=1 ‡ ( − 1)! Q −1<br />

=1 ‡ <br />

(8.105)<br />

This expression gives the correct value of ‡ ( − 0 )=1at = 0 .


8.4 The specific unimolecular reaction rate constants () 198<br />

b) Density of states:<br />

() = ( + () ) −1<br />

( − 1)! Q <br />

=1 <br />

(8.106)<br />

c) Specific rate constant:<br />

³<br />

() = 1 − 0 + ‡ ( − 0 ) ‡ ´−1<br />

() + <br />

()( − 1)! Q −<br />

−1<br />

=1 ‡ <br />

(4) For energies sufficiently above 0 , we can use the expressions<br />

and<br />

to obtain<br />

y<br />

‡ ( − 0 )=<br />

() = ‡ ( − 0 )<br />

()<br />

() =<br />

() =<br />

=<br />

³<br />

− 0 + ‡ ´−1<br />

<br />

( − 1)! Q −1<br />

=1 ‡ <br />

( + ) −1<br />

( − 1)! Q <br />

=1 <br />

³<br />

‡ (0) ‡ ´−1<br />

<br />

()( − 1)! Q −1<br />

=1 ‡ <br />

(8.107)<br />

(8.108)<br />

(8.109)<br />

³<br />

Q <br />

=1 − 0 + ‡ ´−1<br />

<br />

<br />

Q −1<br />

=1 ‡ <br />

( + ) −1 (8.110)<br />

Q <br />

=1 <br />

Q −1<br />

=1 ‡ <br />

The resemblence to the Kassel formula is obvious.<br />

Ã<br />

− 0 + ‡ <br />

+ <br />

! −1<br />

(8.111)<br />

8.4.3 The SACM and VRRKM model<br />

More advanced models allow for the tightening of the TS with increasing (variational<br />

RRKM theory = VRRKM theory) or even for individual adiabatic channel potential<br />

curves (statistical adiabatic channel model = SACM). The latter model has to be used<br />

for reactions with loose transition states.


8.4 The specific unimolecular reaction rate constants () 199<br />

8.4.4 Experimental results<br />

I<br />

Figure 8.19: Experimental methods for the preparation of highly vibrationally excited<br />

molecules with definedexcitationenergy.<br />

I<br />

Figure 8.20: Comparison of theoretically predicted specificrateconstants () for unimolecular<br />

isomerization of cycloheptatrienes with directly measured experimental data.


8.5 Collisional energy transfer* 200<br />

8.5 Collisional energy transfer*<br />

Skipped for lack of time.


8.6 Recombination reactions 201<br />

8.6 Recombination reactions<br />

See homework assignment.


8.6 Recombination reactions 202<br />

8.7 References<br />

Lindemann 1922 F. A. Lindemann, Trans. Faraday Soc. 17, 598 (1922).<br />

Hinshelwood 1927 C. N. Hinshelwood, Proc. Roy. Soc. A113, 230 (1927).<br />

Kassel 1928a J. Phys. Chem. 32, 225 (1928).<br />

Kassel1928b J. Phys. Chem. 32, 1065 (1928).<br />

Kassel 1932 L. S. Kassel, The <strong>Kinetics</strong> of Homogeneous Gas Reactions, Chem.Catalog,<br />

New York, 1932.<br />

Rice 1927 O. K. Rice, H. C. Ramsperger, J. Am. Chem. Soc. 49, 1617 (1927).<br />

Rice 1928 O. K. Rice, H. C. Ramsperger, J. Am. Chem. Soc. 50, 617 (1928).<br />

Troe1979 J.Troe,J.Phys.Chem.83, 114 (1979).<br />

Troe1994 J. Chem. Soc. Faraday Trans. 90, 2303 (1994).


8.7 References 203<br />

9. Explosions<br />

9.1 Chain reactions and chain explosions<br />

9.1.1 Chain reactions without branching<br />

I<br />

Reaction scheme:<br />

(1) A → X chain initiation<br />

(2) X+B→ P+X chain propagation<br />

(4) X → A chain termination<br />

I<br />

Rate laws:<br />

[X]<br />

<br />

= 1 [A] − 4 [X] (9.1)<br />

[P]<br />

<br />

= 2 [B] [X] (9.2)<br />

As we see, reaction (2) does not appear in the rate law because → !<br />

I Solutions of Eq. 9.1 for two limiting cases (cf. Fig. 9.1):<br />

(1) Solution for → 0 and [A] ≈ const by substitution, using<br />

= 1 [A] − 4 [X]<br />

y<br />

y<br />

y<br />

y<br />

=0:<br />

y<br />

<br />

[X] = − 5 y [X] = − <br />

4<br />

− 1 4<br />

<br />

= y <br />

= − 4<br />

= − 4 = 1 [A] − 4 [X] (9.3)<br />

[X] = 1<br />

4<br />

[A] − − 4<br />

(9.4)<br />

[X] = 0 y = 1<br />

[]<br />

4<br />

(9.5)<br />

[X] = 1 [A]<br />

³ ´<br />

1 − − 4<br />

4<br />

(9.6)


9.1 Chain reactions and chain explosions 204<br />

(2) Solution with steady state assumption for [X] at À (i.e, after the induction<br />

time ):<br />

[X]<br />

≈ 0 (9.7)<br />

<br />

y<br />

[X] <br />

= 1<br />

[A] (9.8)<br />

4<br />

y Considering the free radical concentration profile [X ()], no desasters are happening.<br />

I<br />

Figure 9.1: Evolution of the free radical concentration in a normal chain reaction.<br />

9.1.2 Chain reactions with branching<br />

I<br />

Reaction scheme:<br />

(1) A → X chain initiation<br />

(3) X+B→ P+ X chain branching<br />

= branching factor<br />

(4) X → A chain termination<br />

I<br />

Most important example for a chain branching reaction:<br />

H+O 2 → OH + O (9.9)<br />

This reaction is responsible for flame propagation in all hydrocarbon flames: Since H is<br />

small, it has the highest diffusion coefficient and diffuses rapidly into the unburnt gases.


9.1 Chain reactions and chain explosions 205<br />

I<br />

Rate laws:<br />

[X]<br />

<br />

[P]<br />

<br />

= 1 [A] + ( − 1) 3 [X] [B] − 4 [X] (9.10)<br />

= 3 [B] [X] (9.11)<br />

I<br />

Solutions for three limiting cases:<br />

(1) =1or 4 3 ( − 1) [B]:<br />

[X]<br />

≈ 0 (9.12)<br />

<br />

y<br />

1 [A]<br />

[X] =<br />

(9.13)<br />

4 − ( − 1) 3 [B]<br />

y<br />

[P]<br />

= 3 [B] [X] = (9.14)<br />

<br />

The overall reaction is stable, because the radical concentration is stable. The<br />

product formation rate is final (constant, well behaved).<br />

(2) 4 = 3 ( − 1) [B]: Atshorttimes,[A] [B] ≈ const y<br />

[X]<br />

= 1 [A] = const (9.15)<br />

<br />

y<br />

[X] →∞ (9.16)<br />

The overall reaction turns unstable, because the radical concentration goes to<br />

infinity and thus the product formation rate too, leading to chain explosion!<br />

(3) 4 3 ( − 1) [B]: Atshorttimes,[A] [B] ≈ const y<br />

[X]<br />

+ 4 [X] − 3 ( − 1) [X] [B] = 1 [A]<br />

<br />

(9.17)<br />

[X]<br />

+[X]( 4 − 3 ( − 1) [B]) = 1 [A]<br />

<br />

(9.18)<br />

y<br />

1 [A]<br />

³<br />

´<br />

[X] =<br />

× exp [( − 1) 3 [B] ] − 1 (9.19)<br />

( − 1) 3 [B] − 4<br />

The overall reaction turns unstable, because the radical concentration increases<br />

exponentially, leading to an exponentially growing product formation rate, and<br />

therefore chain explosion!


9.1 Chain reactions and chain explosions 206<br />

I<br />

Figure 9.2: Evolutionofthefreeradicalconcentrationsinachainreactionwithbranching<br />

(three limiting cases).


9.2 Heat explosions 207<br />

9.2 Heat explosions<br />

We consider an exothermic bimolecular reaction in a closed reactor<br />

A+B→ C+D ∆ ª 0 (9.20)<br />

and look at the radial temperature profile in the reactor.<br />

I<br />

Figure 9.3: Temperature profilesinastirredandinanunstirredreactor.<br />

9.2.1 Semenov theory for “stirred reactors” (T =const)<br />

I Heat production in an exothermic reaction (∆ H ª < 0):<br />

• Energy released per volume:<br />

∆ ∆ ª £ ¤ kJ cm<br />

3<br />

(9.21)<br />

• Reaction rate:<br />

• heat release:<br />

<br />

<br />

= − [][] (9.22)<br />

<br />

<br />

= ∆ ª <br />

<br />

= ∆ ª 0 − <br />

| {z }<br />

| {z }<br />

Arrhenius rate constant in pressure units<br />

(9.23)<br />

(9.24)


9.2 Heat explosions 208<br />

• positive feedback effect:<br />

reaction rate increases<br />

y temperature ↑<br />

y reaction rate ↑<br />

y heat explosion !!<br />

I<br />

Figure 9.4: Heat explosions.<br />

I Heat removal: At constant temperature in the stirred reactor, we use Newton’s law<br />

<br />

= ( − 0 ) (9.25)<br />

where<br />

: heat conduction coefficient (depends on material)<br />

: reactor surface area<br />

: temperature<br />

0 : wall temperature<br />

(9.26)<br />

As we see, the heat removal increases (becomes steeper), when we increas the reactor<br />

surface area (at constant reactor volume).<br />

I<br />

Stability limits:<br />

(1) 0 = 1 : Heat removal wins over heat production y negative feedback y stable.<br />

(2) 0 = 2 : Heat removal wins over heat production y negative feedback y stable.<br />

(3) 0 = 3 : Heat production wins over heat removal y positive feedback y heat<br />

explosion, but the system should never reach point [3].<br />

(4) 0 = 4 : tangent y stability limit, the system runs away upon the smallest<br />

perturbation!<br />

(5) 0 = 5 : always unstable!


9.2 Heat explosions 209<br />

I<br />

Conclusions:<br />

• The quantity that determines whether the system is stable or not is the wall<br />

temperature 0 ,because 0 determines the initial point of the heat removal line.<br />

• Smaller surface-to-volume ratio is dangerous, because smaller surface at a given<br />

volume leads to a smaller slope of the heat removal line y instability!<br />

• Security measure: internal cooling system (cold water pipes, heat exchanger).<br />

• Emergency measure: when cooling system fails, the reaction mixture should be<br />

strongly diluted by addition of solvent immediately!<br />

I Quantitative estimation of the stability limits:* At point (4) in Fig. 9.4, we have:<br />

<br />

= <br />

(9.27)<br />

<br />

<br />

<br />

y ( − 0 )= ∆ ª 0 − (9.28)<br />

= <br />

<br />

<br />

(9.29)<br />

y = ∆ ª 0 − <br />

<br />

2 (9.30)<br />

Eq<br />

Eq y 2<br />

<br />

=( − 0 ) (9.31)<br />

y 2 − <br />

+ <br />

y = <br />

2 ± s µ<br />

2<br />

0 =0 (9.32)<br />

2<br />

− <br />

0 (9.33)<br />

Solution is: ⎧<br />

⎪⎨<br />

= 1 ³<br />

+ p ´ ⎫<br />

( 2 2<br />

− 4 0 )<br />

⎪⎬<br />

⎪ ⎩ = 1 ³<br />

− p ´<br />

( 2 2<br />

− 4 0 )<br />

⎪⎭<br />

This can be simplified:<br />

(9.34)<br />

(1) For practically important reactions, we have:<br />

À 0 y 0 ¿ <br />

<br />

(9.35)<br />

(2) We also can exclude unreasonably hight temperatures y only the −-sign before<br />

∗ s<br />

= µ 2<br />

<br />

2 − − <br />

2 0 (9.36)<br />

= <br />

2 − r<br />

<br />

1 − 4 0<br />

(9.37)<br />

2 <br />

√ counts: y<br />

= <br />

2<br />

à r<br />

1 −<br />

!<br />

1 − 4 0<br />

<br />

(9.38)


9.2 Heat explosions 210<br />

(3) We can expand the √ :<br />

= 4 0<br />

; 1; (9.39)<br />

<br />

y √ 1 − =1− 1 2 − 1<br />

2 · 4 2 − (9.40)<br />

y ∗ = µ<br />

<br />

1 − 1+ 2 0<br />

+ 16 · <br />

2 0<br />

2<br />

2 2 · 4 · ( ) 2 + (9.41)<br />

y ∗ = 0 + 2 0<br />

<br />

(9.42)<br />

y ∗ − 0 = 2 0<br />

<br />

(9.43)<br />

(4) Insert into in (I) and power expansion:<br />

· ·<br />

(5) Explosion limit:<br />

2<br />

0<br />

2<br />

· exp<br />

µ <br />

2<br />

0<br />

<br />

∙<br />

− ¸<br />

<br />

=<br />

0<br />

⎡<br />

<br />

= · ∆ ª · 0 · exp ⎣−<br />

´ ⎦ · · (9.44)<br />

0<br />

³1+ 0<br />

<br />

∙<br />

= · ∆ ª · 0 · · exp − ¸<br />

<br />

· 2 · · (9.45)<br />

0<br />

· · <br />

· ∆ ª · · · 0 · · <br />

= (9.46)<br />

⎤<br />

y<br />

∝ · <br />

(9.47)<br />

Power expansion of the exponent in step (4):<br />

⎡<br />

⎤<br />

∙<br />

<br />

exp ⎣−<br />

´ ⎦ =exp −<br />

0<br />

³1+ 1 ¸<br />

1<br />

0<br />

1+<br />

<br />

∙<br />

=exp − 1 <br />

≈ exp<br />

∙− 1 (1 − )¸<br />

<br />

∙<br />

=exp 1 − 1 ¸<br />

<br />

∙<br />

= 1 · exp − 1 ¸<br />

<br />

∙<br />

= · exp − ¸<br />

<br />

0<br />

¡<br />

1 − + 2 ¢¸<br />

= 0<br />

<br />

¿ 1(9.48)<br />

(9.49)<br />

(9.50)<br />

(9.51)<br />

(9.52)<br />

(9.53)


9.2 Heat explosions 211<br />

I Figure 9.5:<br />

9.2.2 Frank<strong>—</strong>Kamenetskii theory for “unstirred reactors”<br />

I<br />

Figure 9.6: Temperature profilesinanunstirredreactor.<br />

Boundary effects are neglected in the Figure. In reality, there is a smooth transition in<br />

the boundary layer.


9.2 Heat explosions 212<br />

9.2.3 Explosion limits<br />

I<br />

Figure 9.7: Explosion limits in the H 2 /O 2 system.<br />

9.2.4 Growth curves and catastrophy theory<br />

I<br />

Exponential growth:<br />

∝ · y <br />

<br />

= (9.54)<br />

I<br />

Hyperbolic growth:<br />

∝<br />

1<br />

( − ) y <br />

= 1+ mit 1+1 (9.55)<br />

I<br />

Heat explosion:<br />

<br />

∝ quadratic, i.e. we get hyperbolic growth ! (9.56)


9.2 Heat explosions 213<br />

I<br />

Figure 9.8: Growth curves.


9.2 Heat explosions 214<br />

10. Catalysis<br />

The net rate of a chemical reaction may be affected by the presence of a third species<br />

even though that species is neither consumed nor produced by the reaction. If the rate<br />

increases, we speak of catalysis. If the rate decreases, we speak of inhibition.<br />

In general, we have to distinguish<br />

• homogeneous catalysis,<br />

• heterogeoues catalysis.<br />

10.1 <strong>Kinetics</strong> of enzyme catalyzed reactions (homogeneous catalysis)<br />

The mechanism describing the homogeneous catalytical activity of enzymes has first<br />

been elucidated by Michaelis and Menten (1914).<br />

I<br />

Experimental observations on enzyme catalysis:<br />

(1) For constant substrate concentration [S] 0<br />

, the reaction rate is proportional to<br />

the concentration of the enzyme [E] 0<br />

∝ [E] 0<br />

for [S] 0<br />

= const (10.1)<br />

(2) At an applied enzyme concentration [E] 0<br />

,thereactionrate first increases with<br />

increasing substrate concentrations [S] 0<br />

andthenreachessaturation,asisshown<br />

in Fig. 10.1.<br />

I<br />

Figure 10.1: Enzyme catalysis: Michelis-Menten kinetics.


10.1 <strong>Kinetics</strong> of enzyme catalyzed reactions (homogeneous catalysis) 215<br />

10.1.1 The Michaelis-Menten mechanism<br />

Basic reaction scheme:<br />

E+S 1<br />

À ES (10.2)<br />

−1<br />

ES 2<br />

→ E+P (10.3)<br />

Typical values of 2 :<br />

2 ≈ 10 2 10 3 s −1 (in some cases up to 10 6 s −1 ) (10.4)<br />

Reaction rate:<br />

[P]<br />

<br />

= − [S]<br />

<br />

= 2 [ES] (10.5)<br />

[ES]<br />

<br />

= 1 [E] [S] − −1 [ES] − 2 [ES] (10.6)<br />

≈ 0 (10.7)<br />

Mass balance:<br />

Steady state ES concentration:<br />

[E] 0<br />

= [E]+[ES] (10.8)<br />

y [E] = [E] 0<br />

− [ES] (10.9)<br />

[S] 0<br />

= [S]+[ES]+[P] (10.10)<br />

y [S] = [S] 0<br />

− [ES] − [P] (10.11)<br />

[ES]<br />

<br />

= 1 ([E] 0<br />

− [ES]) ([S] 0<br />

− [ES] − [P]) − ( −1 + 2 )[ES] (10.12)<br />

≈ 0 (10.13)<br />

Approximations:<br />

y<br />

[ES]<br />

<br />

[E] 0<br />

¿ [S] 0<br />

(10.14)<br />

[ES] ¿ [S] 0<br />

(10.15)<br />

[P] ≈ 0 for early times (10.16)<br />

= 1 [E] 0<br />

[S] 0<br />

− 1 [S] 0<br />

[ES] − ( −1 + 2 )[ES] (10.17)<br />

≈ 0 (10.18)<br />

Steady state ES concentration:<br />

[ES] <br />

=<br />

1 [E] 0<br />

[S] 0<br />

1 [S] 0<br />

+ −1 + 2<br />

(10.19)


10.1 <strong>Kinetics</strong> of enzyme catalyzed reactions (homogeneous catalysis) 216<br />

Resulting rate expression:<br />

y<br />

= [P] = 2 [ES] (10.20)<br />

<br />

1 2 [E]<br />

=<br />

0<br />

[S] 0<br />

(10.21)<br />

1 [S] 0<br />

+ −1 + 2<br />

2 [E]<br />

=<br />

0<br />

1+ (10.22)<br />

−1 + 2<br />

1 [S] 0<br />

=<br />

with the Michaelis-Menten constant<br />

2 [E] 0<br />

1+ [S] 0<br />

(10.23)<br />

= −1 + 2<br />

1<br />

(10.24)<br />

Limiting cases:<br />

(1) [S] 0<br />

→∞y <br />

[S] 0<br />

¿ 1: y<br />

= [P]<br />

<br />

= 2 [E] 0<br />

(10.25)<br />

(2) [S] 0<br />

→ 0 y <br />

[S] 0<br />

À 1: y<br />

= [P]<br />

<br />

= 2 [E] 0<br />

[S] 0<br />

<br />

(10.26)<br />

Determination of 2 and from a Lineweaver-Burk plot (Fig. 10.2):<br />

1<br />

= 1 +<br />

1<br />

(10.27)<br />

2 [E] 0<br />

2 [E] 0<br />

[S] 0<br />

⇒ Aplotof −1 vs. [S] −1 0<br />

should give a straight line with intercept ( 2 [E] 0<br />

) −1 and<br />

slope 2 [E] 0<br />

. y<br />

slope<br />

intercept = = −1 + 2<br />

(10.28)<br />

1


10.1 <strong>Kinetics</strong> of enzyme catalyzed reactions (homogeneous catalysis) 217<br />

I<br />

Figure 10.2: Lineweaver-Burk plot.<br />

10.1.2 Inhibition of enzymes:<br />

Extended reaction mechanism:<br />

E+S 1<br />

À ES (10.29)<br />

−1<br />

ES 2<br />

→ E+P (10.30)<br />

E+I 3<br />

À EI (10.31)<br />

−3<br />

The EI complex is catalytically inactive, since I blockstheactivesiteofE.<br />

Inhibition constant:<br />

= [EI]<br />

[E] [I] = 3<br />

−3<br />

(10.32)<br />

I<br />

Reaction rate (Fig. 10.2):<br />

y<br />

[ES]<br />

<br />

= ≈ 0 (10.33)<br />

1<br />

= 1 +(1+ [I]) ×<br />

1<br />

(10.34)<br />

2 [E] 0<br />

2 [E] 0<br />

[S] 0<br />

y The slope of the plot of −1 vs. [S] −1 0<br />

goes up (see Fig. 10.2).<br />

Exercise 10.1: Derive Eq. 10.34 with a reasonable assumption for the concentrations<br />

of I. ¤


10.2 <strong>Kinetics</strong> of heterogeneous reactions (surface reactions) 218<br />

10.2 <strong>Kinetics</strong> of heterogeneous reactions (surface reactions)<br />

I Examples for heterogeneous reactions (gas-solid interface): Many reactions,<br />

which are very slow in the gas phase, take place as heterogeneous reactions in the<br />

presence of a heterogeneous catalyst:<br />

• catalytic hydrogenation<br />

• NH 3 synthesis (Haber-Bosch)<br />

• catalytic exhaust gas treatment (internal combustion engines)<br />

• catalytic DeNO processes<br />

• CO oxidation<br />

• oxidative coupling reactions of CH 4 (oxide catalysts)<br />

• atmospheric reactions (at gas-solid and gas-liquid interface)<br />

10.2.1 Reaction steps in heterogeneous catalysis<br />

Heterogeneous reactions take place via a series of steps:<br />

• Diffusion in gas (or liquid) phase to catalyst surface.<br />

• Adsorption:<br />

<strong>—</strong> “physical” adsorption (“physisorption”),<br />

<strong>—</strong> “chemical” adsorption (“chemisorption”).<br />

• Diffusiononthesurface.<br />

• <strong>Chemical</strong> reaction on the surface.<br />

• Desorption.<br />

• Diffusion away from surface.<br />

10.2.2 Physisorption and chemisorption<br />

We distinguish between “physical” adsorption (“physisorption”) and “chemical” adsorption<br />

(“chemisorption”) on a surface based upon the adsorption energies (surface binding<br />

energies). ⇒ Fig. 10.3.<br />

• Physisorption:<br />

<strong>—</strong> van-der-Waals type interaction,<br />

<strong>—</strong> typical bond energy: ∆ ≤−40 kJ mol −1 ,


10.2 <strong>Kinetics</strong> of heterogeneous reactions (surface reactions) 219<br />

<strong>—</strong> example: Ar on a metal or oxide surface.<br />

• Chemisorption:<br />

<strong>—</strong> formation of a chemical bond between adsorbate and surface.<br />

<strong>—</strong> typical bond energy: ∆ ≤−100 − 300 kJ mol −1 ,<br />

<strong>—</strong> example: CO on Pd:<br />

O<br />

||<br />

C<br />

|<br />

− Pd − Pd −<br />

(10.35)<br />

∗ chemical bond by overlap of orbital of the C atom with free bands of<br />

Pd,<br />

∗ back donation from occupied bands into ∗ orbital of the CO substrate.<br />

∗ Experiment shows that (Pd − C=O)≈ 236 cm −1 .<br />

I<br />

Figure 10.3: Potential curves for physisorption and chemisorption.<br />

10.2.3 The Langmuir adsorption isotherm<br />

I<br />

Reaction scheme:<br />

A() +<br />

−<br />

|<br />

S − 1<br />

À<br />

−1<br />

−<br />

A<br />

|<br />

S − (10.36)<br />

− | S− = “active surface site” (10.37)


10.2 <strong>Kinetics</strong> of heterogeneous reactions (surface reactions) 220<br />

I Fraction of occupied active surface sites Θ:<br />

• Adsorption rate:<br />

• Desorption rate:<br />

• Steady state:<br />

y<br />

y<br />

1 × (1 − Θ) × (10.38)<br />

−1 × Θ (10.39)<br />

1 (1 − Θ) = −1 Θ (10.40)<br />

1 =( −1 + 1 ) Θ (10.41)<br />

I Langmuir isotherm (Fig. 10.4):<br />

Θ =<br />

1 <br />

−1 + 1 =<br />

<br />

1+ <br />

(10.42)<br />

with<br />

= 1<br />

−1<br />

(10.43)<br />

I<br />

Figure 10.4: Langmuir adsorption isotherm.


10.2 <strong>Kinetics</strong> of heterogeneous reactions (surface reactions) 221<br />

I<br />

Determination of V , V ∞ , and the adsorption enthalpy ∆H : At a series<br />

of temperatures , we need to measure Θ ( )= ( )<br />

∞ ( ) as function of by<br />

pressure/volume measurements in an ultrahigh vacuum apparatus.<br />

• From Eq. 10.42:<br />

• Multiplication of l.h.s. and r.h.s. with<br />

• Aplotof<br />

<br />

<br />

=<br />

1<br />

Θ = <br />

∞<br />

=1+ 1<br />

<br />

1<br />

∞<br />

<br />

:<br />

<br />

∞<br />

+ <br />

∞<br />

<br />

(10.44)<br />

= + (10.45)<br />

<br />

<br />

vs. gives a straight line (see the inset in Fig. 10.4) with slope<br />

=( ∞<br />

) −1 (10.46)<br />

and intercept<br />

=( ∞<br />

) −1 (10.47)<br />

•<br />

• van’t Hoff plot:<br />

=<br />

slope<br />

intercept = <br />

ln <br />

(1 ) = −∆ <br />

<br />

(10.48)<br />

(10.49)<br />

I Desorption rate: If ∆ 0 0 we can describe the desorption rate by TST (see Figs.<br />

10.3 und 10.6).<br />

−1 = µ<br />

‡<br />

exp − ∆ <br />

0<br />

(10.50)<br />

<br />

The ’s are partition functions per unit area.


10.2 <strong>Kinetics</strong> of heterogeneous reactions (surface reactions) 222<br />

10.2.4 Surface Diffusion<br />

I<br />

Figure 10.5: Surface Diffusion.<br />

I<br />

Diffusion is an activated process:<br />

µ<br />

= 0 exp − ∆ <br />

<br />

<br />

∆ = energy barrier of jump of adsorbate from one site to another.<br />

(10.51)<br />

I<br />

Fick’s 2nd law:<br />

( )<br />

<br />

= × 2 ( )<br />

2 (10.52)<br />

I<br />

Einstein’s diffusion law:<br />

• 1D:<br />

∆ 2 =2 (10.53)<br />

• 2D:<br />

∆ 2 =4 (10.54)


10.2 <strong>Kinetics</strong> of heterogeneous reactions (surface reactions) 223<br />

10.2.5 Types of heterogeneous reactions<br />

a) Dissociative Adsorption<br />

I<br />

Scheme:<br />

A 2 () +<br />

−<br />

|<br />

S −<br />

|<br />

S − 2<br />

À<br />

−2<br />

−<br />

A<br />

|<br />

S −<br />

A<br />

|<br />

S − (10.55)<br />

I<br />

Adsorption isotherm:<br />

• Adsorption rate:<br />

• Desorption rate:<br />

y<br />

with<br />

Θ =<br />

2 2 (1 − Θ) 2 (10.56)<br />

−2 Θ 2 (10.57)<br />

( 2<br />

× ) 12<br />

1+( 2 × ) 12 (10.58)<br />

= 2<br />

−2<br />

(10.59)<br />

I<br />

Figure 10.6: Potential energy curves for dissociative adsorption.


10.2 <strong>Kinetics</strong> of heterogeneous reactions (surface reactions) 224<br />

b) Surface catalyzed unimolecular dissociation<br />

I<br />

Scheme:<br />

A() +<br />

−<br />

|<br />

S − 3<br />

À<br />

−3<br />

−<br />

A<br />

|<br />

S − 4<br />

→ Products (10.60)<br />

I Example: Decomposition of PH 3 on a W metal surface.<br />

I<br />

Decomposition rate and two limiting cases:<br />

[P]<br />

<br />

= 4 Θ = 4 × A<br />

1+ A<br />

(10.61)<br />

(1) A ¿ 1: The reaction becomes first-order in A .<br />

[P]<br />

<br />

= 4 A (10.62)<br />

(2) A À 1: The reaction becomes zeroth order, i.e., independent of A !<br />

[P]<br />

<br />

= 4 (10.63)<br />

c) Bimolecular reactions<br />

I<br />

Scheme:<br />

A() +<br />

−<br />

|<br />

S − <br />

À<br />

−<br />

−<br />

A<br />

|<br />

S − (10.64)<br />

−<br />

A<br />

|<br />

S − + −<br />

B() +<br />

B<br />

−<br />

|<br />

S − 5<br />

→<br />

|<br />

S −<br />

−<br />

<br />

À<br />

−<br />

A − B<br />

−<br />

B<br />

|<br />

S − (10.65)<br />

|<br />

S − S − 6<br />

→ Products (10.66)<br />

I Reaction rate: Assuming that A and B compete for the same active sites, we have<br />

(1 − Θ − Θ )= − Θ (10.67)<br />

y<br />

[P]<br />

<br />

(1 − Θ − Θ )= − Θ (10.68)<br />

= 5 Θ Θ =<br />

5 <br />

(1 + + ) 2 (10.69)


10.2 <strong>Kinetics</strong> of heterogeneous reactions (surface reactions) 225<br />

I Limiting cases: It is often the case that one species is adsorbed much more strongly<br />

than the other.<br />

• Assume, for instance, that ¿ . Then,<br />

[P]<br />

<br />

= 5 <br />

(1 + ) 2 (10.70)<br />

(1) À 1:<br />

[P]<br />

<br />

= 5 <br />

<br />

(10.71)<br />

The reaction is inhibited by B, which is strongly adsorbed: At high ,all<br />

active sites are blocked by B.<br />

(2) ¿ 1:<br />

[P]<br />

= 5 (10.72)<br />

<br />

The reaction is truly second order.<br />

⇒ We can see from these two cases that, as increases, the rate [P]<br />

<br />

increases, reaches a maximum, and then decreases again.<br />

first<br />

d) Example 1: Heterogeneous oxidation of CO<br />

I<br />

Figure 10.7: Heterogeneous oxidation of CO on Pt. (a) Langmuir-Hinshelwood and<br />

Eley-Rideal mechanisms, (b) potential energy diagram (values in kJ mol −1 ).


10.2 <strong>Kinetics</strong> of heterogeneous reactions (surface reactions) 226<br />

e) Example 2: Haber-Bosch NH 3 synthesis<br />

I Figure 10.8: Potential energy diagram for the catalytic reaction N 2 + 3H 2 → 2NH 3<br />

on a Fe 3 O 4 catalyst according to Ertl and co-workers (1982). Energies in kJ mol −1 .<br />

f) How can we control and accelerate heterogeneous reactions?<br />

The answer depends on which step is rate determining:<br />

• If diffusion to/from the surface is the rate determining step, we can accelerate<br />

the reaction by stirring<br />

• If the adsorption step is rate determining (because of an activation energy), we<br />

can accelerate the reaction by increasing and and, in particular, by increasing<br />

surface area.


10.2 <strong>Kinetics</strong> of heterogeneous reactions (surface reactions) 227<br />

11. Reactions in solution<br />

“Everyone has problems - chemists have solutions.”<br />

I<br />

Comparison of gas and liquid phase reactions (@ T = 298 K and p =1bar):<br />

• Molecules in the gas phase occupy ≈ 02 % of the total volume, while molecules<br />

in the liquid phase occupy ≈ 50 % of the total volume.<br />

• Molecules in liquid are in permanent contact with each other, and undergo permanent<br />

collisions with each other.<br />

• The crossing of a potential barrier in a reaction is not continuous as in gas phase<br />

but affected by multiple collisions during the crossing.<br />

• As a result of the permanent collisions, we can expect the reactant molecules to<br />

be in thermodynamic equilibrium even above 0 . Thus TST should be applicable.<br />

• However, it is very difficult to evaluate the partition functions in liquids. Hence,<br />

one usually employs the thermodynamic version TST.<br />

• Solvent molecules act as an efficient heat bath.<br />

I Observations on liquid phase reactions: Experience has shown that reactions in<br />

liquids have similar rates as in the gas phase, except if<br />

• the reaction occurs with the solvent,<br />

• there are strong interactions of the reactant molecules with the solvent (e.g., ionic<br />

reactions; ⇒ solvent shells),<br />

• the reactions are diffusion controlled.<br />

11.1 Qualitative model of liquid phase reactions<br />

In contrast to gas phase reactions for which we have advanced theories (⇒ transition<br />

state theory, unimolecular rate theory), the theory for liquid phase reactions is much<br />

less well developed. However, we can understand the required basic reaction steps as<br />

sketched in Fig. 11.1 and a number of limiting cases resulting from these steps.


11.1 Qualitative model of liquid phase reactions 228<br />

I<br />

Figure 11.1: Structure and dynamics of solutions.<br />

I Formal kinetics: We distinguish between the formation of the encounter complex in<br />

the solvent cage (by diffusion) and the actual reaction:<br />

+ <br />

<br />

À<br />

−<br />

{}<br />

<br />

→ (11.1)<br />

y<br />

y<br />

with<br />

[{}]<br />

<br />

[ ]<br />

<br />

= [][] − − [{}] − [{}] ≈ 0 (11.2)<br />

[{}] =<br />

<br />

− + <br />

[][] (11.3)<br />

= [{}] = <br />

− + <br />

[][] = [][] (11.4)<br />

=<br />

<br />

− + <br />

(11.5)<br />

I<br />

Rate constant for liquid phase reactions:<br />

=<br />

<br />

− + <br />

(11.6)


11.1 Qualitative model of liquid phase reactions 229<br />

I<br />

Two limiting cases:<br />

(1) Diffusion control: À − y<br />

= (11.7)<br />

This case is observed when =0and the diffusion coefficient is small.<br />

We shall find below that<br />

=4 (11.8)<br />

(2) Reaction (or activation) control: ¿ − y<br />

= <br />

−<br />

= {} (11.9)<br />

In this case, the reaction is said to be activation controlled because usually it has<br />

a sizable .<br />

can be described by TST:<br />

= <br />

∆+ −∆+ <br />

(11.10)


11.2 Diffusion-controlled reactions 230<br />

11.2 Diffusion-controlled reactions<br />

11.2.1 Experimental observations<br />

The transition to diffusion control can be observed in the gas phase by studying the<br />

kinetics in highly compressed gases (up to =1000bar) or, even better, in supercritical<br />

media. At ≈ 1000 bar, the gas density reaches the density of the liquid phase.<br />

I Smoluchowsky-Kramers limit: At high densities, the rate constant becomes<br />

∝ ∝ 1 <br />

(11.11)<br />

I<br />

Figure 11.2: Diffusion control of a unimolecular reaction: At very high pressures, the<br />

fragments recombine before they become separated by diffusion y the reaction becomes<br />

diffusion controlled.<br />

11.2.2 Derivation of k <br />

In order to derive a theoretical expression for , we consider the distribution of the<br />

molecules around a molecule inthecenterofasphereasshowninFig.11.3.


11.2 Diffusion-controlled reactions 231<br />

I<br />

Figure 11.3: Radial distribution of reactants in a diffusion-controlled reaction.<br />

(1) We start by considering the boundary conditions for the concentration [] (cf.<br />

Fig. 11.3):<br />

• At a distance →∞, the concentration of is identical to the mean<br />

concentration in the solution, i.e.,<br />

→∞y [] <br />

=[] =∞<br />

=[] (11.12)<br />

• At the distance = for a diffusion controlled reaction, the concentration<br />

of vanishes because of its fast reaction with , i.e.,<br />

→ 0 y [] =0<br />

=0 (11.13)<br />

(2) The result is a concentration gradient [] <br />

around , which gives rise to a net<br />

<br />

flux ←− by diffusion as described by Fick’s first law:<br />

= × × [] <br />

(11.14)<br />

<br />

We define this flux, measured in units of molecules/s, in the direction from = ∞<br />

(high concentration) to = (low concentration); thus there is no − sign in<br />

the usually written form of Fick’s law.<br />

is the mean diffusion coefficient, measured in m 2 s,<br />

= + (11.15)


11.2 Diffusion-controlled reactions 232<br />

(3) With =4 2 as the surface of the sphere with radius around , weobtain<br />

= × 4 2 × [] <br />

<br />

(11.16)<br />

(4) To determine the concentration profile [] <br />

, we integrate over [] <br />

:<br />

Z []∞ Z ∞<br />

<br />

[] <br />

=<br />

<br />

[] <br />

4 2 <br />

(11.17)<br />

y<br />

∞<br />

[]| ∞ <br />

<br />

= −<br />

4 <br />

¯¯¯¯<br />

(11.18)<br />

y<br />

<br />

[] − [] <br />

= − 0 (11.19)<br />

4 <br />

(5) is found from the boundary conditions that [] <br />

=0at = :<br />

<br />

=4 × [] (11.20)<br />

(6) The concentration gradient of [] is obtained by inserting the above result for <br />

into Eq. 11.19:<br />

[] − [] <br />

= []<br />

(11.21)<br />

<br />

y<br />

³<br />

[] <br />

=[] 1 − ´<br />

<br />

(11.22)<br />

<br />

This expression is plotted in Fig. 11.3.<br />

(7) To determine , we write the rate of the diffusion controlled reaction as<br />

[ ]<br />

= [] (11.23)<br />

<br />

Inserting the expression for from Eq. 11.20, we obtain<br />

[ ]<br />

=4 × [][] (11.24)<br />

<br />

The diffusion-limited rate constant is thus (in molecular units)<br />

=4 (11.25)<br />

(8) Nernst-Einstein relation between diffusion constant and viscosity for spherical particles<br />

with radius :<br />

= <br />

(11.26)<br />

6 <br />

y<br />

∝ <br />

(11.27)<br />

<br />

(9) Keepinmindthatdiffusion is an activated process:<br />

= 0 − <br />

Typical value for H 2 O: ≈ 15 kJ mol −1 .<br />

(11.28)


11.2 Diffusion-controlled reactions 233<br />

I<br />

Typical values:<br />

y Typical magnitude of :<br />

≈ 2 × 10 −5 cm 2 s −1 (11.29)<br />

≈ 5 × 10 −8 cm (11.30)<br />

≈ 8 × 10 9 lmol −1 s −1 (11.31)<br />

I<br />

Table 11.1: Diffusion constants of ions in aqueous solution at =298K.<br />

ion (cm 2 s −1 ) ion (cm 2 s −1 )<br />

H + 91 × 10 −5 OH − 52 × 10 −5<br />

Li + 10 × 10 −5 Cl − 20 × 10 −5<br />

Na + 13 × 10 −5 Br − 21 × 10 −5<br />

I<br />

Examples of diffusion controlled reaction in solution:<br />

• radical recombination reactions, e.g.<br />

+ → 2 : =13 × 10 10 lmol −1 s −1 (11.32)<br />

• electronic deactivation (quenching) in solution<br />

• H + transfer reactions<br />

+ + − → 2 : =14 × 10 11 lmol −1 s −1 (11.33)<br />

This reaction is ten times faster than a typical diffusion controlled reaction, because<br />

H + transfer is a concerted process (Grotthus mechanism, see Fig. 11.4)!<br />

I<br />

Figure 11.4: Structures involved in fast H + transfer processes in liquid water.


11.3 Activation controlled reactions 234<br />

11.3 Activation controlled reactions<br />

The rate constant for an activation controlled reaction is<br />

= {} (11.34)<br />

I Estimation of K {} : We can estimate the equilibrium constant {} using simple<br />

arguments based on the coordination number. Accordingly, the concentration of the<br />

encounter complex is<br />

[{}] =[] × (probability of finding next to ) (11.35)<br />

=[] × []<br />

[]<br />

(11.36)<br />

where [] is the solvent concentration and is the coordination number. y<br />

{} = [{}]<br />

[][] = <br />

[]<br />

(11.37)<br />

For H 2 Oat =298K, [] =555moll −1 . Thus, taking for example =8,wefind<br />

that<br />

{} =014 l mol −1 (11.38)<br />

I Application of thermodynamic TST: TSTcanbeappliedintwoways:<br />

(1) Considering the overall reaction, we can write<br />

= <br />

exp ¡ −∆ ‡ ¢ (11.39)<br />

where ∆ ‡ is the overall Gibbs free enthalphy for the reaction.<br />

(2) We can also separate the effects from and {} by writing<br />

´<br />

{} =exp<br />

³−∆ ª {} <br />

(11.40)<br />

and<br />

= <br />

exp ¡ −∆ + ¢ (11.41)<br />

∆ ª {}<br />

is the Gibbs free enthalpy change for the formation of the encounter<br />

complex, and ∆ + is the Gibbs free enthalphy of activation from the encounter<br />

complex. Thus,<br />

= ³<br />

<br />

exp −<br />

³∆ ∆+´ ´<br />

ª<br />

{} + (11.42)<br />

I Pressure dependence of the rate constant: ⇒ See section 7.3.5.


11.4 Electron transfer reactions (Marcus theory) 235<br />

11.4 Electron transfer reactions (Marcus theory)<br />

Electron transfer is what happens in oxidation-reduction reactions. Hence, electron<br />

transfer reactions are extremely important. They are strongly affected by the solvent<br />

because the change of the charge distribution results in a large reorganization energy.<br />

I Figure 11.5:<br />

I Marcus theory: For a simplified derivation of ( ) we refer to Fig. 11.5.<br />

(1) The energy of the donor and acceptor molecules will generally depend on all 3−6<br />

internal coordinates. We consider the dependence on an effective coordinate ,<br />

which describes the minimum energy path from reactant to product (Fig. 11.5).<br />

(2) The exergonicity of the ET reaction ∆ ª is negative for an exergonic reaction<br />

(convention; cf. Fig. 11.5).<br />

(3) For the donor (D) and the acceptor (A), the Gibbs energy profiles () and<br />

() shall be of parabolic form, which we may write as<br />

() = 2 (11.43)<br />

and<br />

() =( − ) 2 + ∆ ª . (11.44)<br />

The origin =0has been placed at the minimum of the donor; the acceptor has<br />

its minimum at = . 50<br />

50 We assume that ∝ 2 , taking any proportionality constant into account by proper rescaling of the<br />

abscissa.


11.4 Electron transfer reactions (Marcus theory) 236<br />

(4) The ET rate constant is described by TST as<br />

= <br />

exp ¡ −∆ + ¢ (11.45)<br />

The Gibbs free energy of activation<br />

∆ + = 2 (11.46)<br />

is the difference from the donor minimum to the TS at the donor/acceptor curve<br />

crossing at point .<br />

(5) In order to relate ∆ + to the potential curves, we define the “reorganization<br />

energy” as the energy that the acceptor would have at the equilibrium<br />

geometry of the donor. This is the energy of the acceptor parabola () at<br />

=0.<br />

If we count from the minimum of the acceptor parabola (), wehave<br />

=(− ) 2 = 2 (11.47)<br />

(6) We now determine the energy of the intersection point of the two parabolas<br />

from the condition<br />

( )= ( ) . (11.48)<br />

Inserting into Eqs. 11.43 and 11.44 yields, and using Eq. 11.47, yields<br />

2 =( − ) 2 + ∆ ª (11.49)<br />

= 2 − 2 + 2 + ∆ ª (11.50)<br />

= 2 − 2 + + ∆ ª (11.51)<br />

y<br />

y<br />

y<br />

y<br />

2 = + ∆ ª (11.52)<br />

∆ + = 2 = ( + ∆ ª ) 2<br />

( )= <br />

<br />

= + ∆ ª<br />

2 <br />

(11.53)<br />

= ( + ∆ ª ) 2<br />

(11.54)<br />

4 <br />

2 4 <br />

Ã<br />

!<br />

exp − ( + ∆ ª ) 2<br />

(11.55)<br />

4 <br />

(7) Limiting cases: We can distinguish between<br />

• −∆ ª : Regular region – the reaction becomes faster, the more<br />

exergonic the reaction becomes.<br />

• −∆ ª = : turning point<br />

• −∆ ª : Inverted region – the reaction becomes slower, the more<br />

exergonic the reaction becomes.


11.4 Electron transfer reactions (Marcus theory) 237<br />

I Conclusion: Equation 11.55 predicts the ET rate constant to increase with increasing<br />

exoergicity (−∆ ª ) of the reaction. However, once −∆ ª , the rate constant<br />

is predicted to decrease again (see Figs. 11.6D and 11.7). The prediction of this<br />

unexpected inverted Marcus regime before any experimental data were available is<br />

the hallmark of Marcus theory.<br />

I Figure 11.6:<br />

I Figure 11.7:


11.5 Reactions of ions in solutions 238<br />

11.5 Reactions of ions in solutions<br />

Reactions of ionic species in liquids are one exception to the rule that reactions in the<br />

liquid phase usually have similar rates as in the gas phase (the other exception being<br />

electron transfer reactions). The reason is that the charged species are very sensitive<br />

to their environment, in particular solvation. Changes of the charge distribution are<br />

accompanied by correspondingly large solvation shell rearrangements.<br />

11.5.1 Effect of the ionic strength of the solution<br />

The main effect arises from the stabilization of shielding every ion in solution by the oppositely<br />

charged ”ionic atmosphere” or “ion cloud” (⇒ Debye-Hückel theory, Appendix<br />

??).<br />

I<br />

Equilibrium constant for<br />

hAB ‡ i<br />

:<br />

‡ = ‡ <br />

<br />

= ‡ <br />

<br />

£<br />

<br />

‡ ¤<br />

[][]<br />

(11.56)<br />

I<br />

Activity coefficient from Debye-Hückel theory:<br />

log = − 2 12<br />

(11.57)<br />

with (for water at =298K, according to Debye-Hückel theory)<br />

=0509 l 12 mol −12 (11.58)<br />

I<br />

Ionic strength:<br />

= 1 X<br />

2 (11.59)<br />

2<br />

<br />

11.5.2 Kinetic salt effect<br />

I<br />

TST rate constant:<br />

[ ]<br />

<br />

= £ ‡ ‡¤ = ‡ <br />

‡ [][] (11.60)<br />

‡ <br />

Defining 0 as the rate constant for the ideal solution, where all =1, this becomes<br />

= 0<br />

<br />

‡ <br />

(11.61)


11.5 Reactions of ions in solutions 239<br />

I<br />

Dependence of the rate constant on the ionic strength:<br />

³<br />

´<br />

log =log 0 − 2 + 2 − ‡ 2<br />

<br />

12<br />

(11.62)<br />

with ‡ = + .<br />

Inserting ‡ = + ,weobtain<br />

log =log 0 − ¡ 2 + 2 − ( + ) 2¢ 12<br />

(11.63)<br />

y<br />

log 0<br />

=+2 12<br />

(11.64)<br />

I<br />

Conclusions:<br />

• If and havethesamesign,then will increase with increasing , because<br />

the repulsion is reduced by the shielding effect of the ion cloud.<br />

• If and have opposite sign, then will decrease with increasing , because<br />

the attraction is reduced by the shielding effect of the ion cloud.<br />

I<br />

Figure 11.8: Kinetic salt effect for reactions of ions in solution.


11.5 Reactions of ions in solutions 240<br />

11.6 References<br />

Eigen 1954a M. Eigen, Disc. Faraday Soc. 17, 194 (1954).<br />

Eigen 1954b M.EigenZ.Physik.Chem.NF1, 176 (1954).


11.6 References 241<br />

12. Photochemical reactions<br />

I<br />

Primary photochemical processes:<br />

• Absorption of a photon (depends on transition dipole moment ¯¯ ¯¯2),<br />

• Intramolecular processes:<br />

<strong>—</strong> fluorescence (≡ spontaneous emission of a photon, FL),<br />

<strong>—</strong> stimulated emission (SE), (⇒ Einstein coefficients)<br />

<strong>—</strong> chemical reaction in the excited state (R),<br />

<strong>—</strong> intramolecular vibrational (energy) redistribution (IVR),<br />

<strong>—</strong> internal conversion (IC),<br />

<strong>—</strong> intersystem crossing (ISC),<br />

<strong>—</strong> phosphorescence (PHOS).<br />

• Intermolecular (i.e., collisional) processes:<br />

<strong>—</strong> electronic deactivation (quenching, Q),<br />

∗ may affect fluorescence as well as phosphorescence,<br />

<strong>—</strong> vibrational energy transfer (VET),<br />

<strong>—</strong> rotational energy transfer (RET),<br />

<strong>—</strong> collision-induced intersystem crossing (CIISC).<br />

12.1 Basic radiative processes in a two-level system; Einstein coefficients<br />

I<br />

Figure 12.1: Absorption, stimulated emission, and spontaneous emission.<br />

2<br />

<br />

= − 1<br />

<br />

= 12 () 1 − 21 () 2 − 21 2 (12.1)


12.2 Fluorescence quenching (Stern-Volmer equation) 242<br />

12.2 Fluorescence quenching (Stern-Volmer equation)<br />

I<br />

Kinetic model:<br />

A + → A ∗ 1<br />

A ∗ → A + <br />

A ∗ +Q → A + Q <br />

(12.2)<br />

If any other radiationless processes can be neglected, we find<br />

[A ∗ ] <br />

= 1 [A]<br />

+ [Q]<br />

(12.3)<br />

I Fluorescence intensity I 0 in the absence of the quencher Q:<br />

0 ∝ [A ∗ ] <br />

= 1 [A] (12.4)<br />

I Fluorescence intensity I in the presence of the quencher Q:<br />

∝ [A ∗ ] <br />

= <br />

1 [A]<br />

+ [Q]<br />

(12.5)<br />

I<br />

Stern-Volmer equation:<br />

y<br />

0<br />

<br />

= 1 [A]<br />

1 [A]<br />

+ [Q]<br />

= + [Q]<br />

<br />

=1+ [Q] (12.6)<br />

0<br />

<br />

=1+ 0 [Q] (12.7)<br />

0 =( ) −1 is the radiative lifetime.<br />

I Conclusion: Aplotof 0 vs. [Q] should give a straight line with intercept 1 and<br />

slope 0 [Q].


12.3 The radiative lifetime 0 and the fluorescence quantum yield Φ 243<br />

12.3 The radiative lifetime τ 0 and the fluorescence quantum yield Φ <br />

I Determination of the radiative lifetime τ 0 : Assuming that there are no competing<br />

nonradiative processes, we can determine the radiative lifetime 0 by<br />

(1) time-resolved measurements of the (exponential) fluorescence decay after pulsed<br />

excitation,<br />

(2) high-resolution measurement of the natural (Lorentzian) linewidth according to<br />

∆ ∆ = ~ (12.8)<br />

y<br />

0 =<br />

1<br />

2 ∆ 0<br />

(12.9)<br />

(3) evaluation from the molar absorption coefficient (Lambert-Beer) using the relationship<br />

between the Einstein coefficients 21 and 21 , 51<br />

21 = 83<br />

3 21 (12.10)<br />

(4) calculate the value of ¯¯ ¯¯2 (⇒ 12 21 21 )fromfirst principles.<br />

I Fluorescence quantum yield: Apart from quenching, the fluorescence is reduced by<br />

other radiationless processes (IC, ISC, R, . . . ). The experimentally measured fluorescence<br />

lifetime is therefore smaller (often much smaller) than the radiative lifetime 0 .<br />

The ratio is the fluorescence quantum yield:<br />

Φ =<br />

<br />

+ + + + [Q] + = 0<br />

(12.11)<br />

51 The actual procedure, which involves integration over the absorption band, corrects for the influence<br />

of the refraction index of the solvent, etc., is called the Strickler-Berg analysis.


12.4 Radiationless processes in photoexcited molecules 244<br />

12.4 Radiationless processes in photoexcited molecules<br />

I<br />

Figure 12.2: Jablonski Diagram.<br />

Elementary Photochemical Processes: Jablonski Diagram<br />

Energy<br />

RXN<br />

(10 -14 .. 10 -6 s)<br />

S 2<br />

T 2<br />

IC<br />

S 2 -S 1 Abs.<br />

T 1<br />

T 2 -T 1 Abs.<br />

IVR<br />

IC<br />

VET* S 1<br />

ISC<br />

(10 -9 ..10 -3 s)<br />

S 0 VET* (10 -11 s)<br />

IVR (10 -13 ..10 -9 s)<br />

Absorption<br />

(10 -18 s)<br />

Fluorescence<br />

(10 -9 ..10 -8 s)<br />

Phosphorescence<br />

(10 -3 ..10 -6 s)<br />

*also known as VR<br />

<strong>Physical</strong> <strong>Chemistry</strong> III: <strong>Chemical</strong> <strong>Kinetics</strong> • © F. Temps, IPC Kiel 253<br />

12.4.1 The Born-Oppenheimer approximation and its breakdown<br />

a) The full Hamiltonian<br />

We consider a non-moving, non-rotating molecule in the laboratory framework. The<br />

molecule is described by the Schrödinger equation (SE)<br />

ˆ(r R) = (r R) (12.12)<br />

or ³<br />

ˆ − ´<br />

(r R) =0 (12.13)<br />

with being the energy eigenvalue associated with the wavefunction (r R) as function<br />

of the electronic (e) coordinates r and the nuclear (N) coordinates R.<br />

The Hamilton operator ˆ appearingintheSEcanbewrittenas<br />

ˆ = ˆ + ˆ = ˆ (r)+ ˆ (R)+ (r R) (12.14)<br />

where<br />

ˆ = − ~2<br />

2 <br />

X <br />

=1<br />

∇ 2 (12.15)<br />

ˆ = − ~2 X <br />

∇ 2 (12.16)<br />

2 <br />

=1


12.4 Radiationless processes in photoexcited molecules 245<br />

and<br />

with<br />

(r R) = (R)+ (r R)+ (r) (12.17)<br />

<br />

<br />

<br />

=+ 2<br />

4 0<br />

X<br />

<br />

X<br />

0 0 =1<br />

= − 2<br />

4 0<br />

<br />

X<br />

=1<br />

X <br />

=1<br />

=+ 2 X X <br />

4 0<br />

0<br />

0<br />

<br />

0 0 0<br />

=1<br />

(12.18)<br />

<br />

<br />

(12.19)<br />

1<br />

(12.20)<br />

Electronic spin, spin-orbit interaction, and nuclear spins are neglected here. Further,<br />

we have left aside the so-called electronic mass and electronic nuclear cross polarization<br />

terms, which appear in the case of a moving molecule as result of the separation of the<br />

center-of-mass.<br />

b) Adiabatic separation of nuclear and electronic motion<br />

We now want to separate the slow motion of the nuclei from the the fast motion of<br />

the electrons. This separation can be made because the fast electrons can be assumed<br />

to very rapidly follow the slow nuclei. The kinetic energy of the nuclei is assumed to be<br />

small compared to the electronic energy.<br />

Thus, for every nuclear configuration R, there is an electronic wavefunction el<br />

(r R)<br />

belonging to the electronic state |i specified by quantum number . This el<br />

(r R)<br />

depends on the nuclear coordinates R, but only very little on the nuclear velocities.<br />

We say that the electrons adiabatically follow the periodic vibrational motion of the<br />

nuclei. This approximation is therefore called adiabatic approximation.<br />

I Ansatz for Ĥ:<br />

with<br />

ˆ 0 = ˆ + (r R) =− ~2<br />

2 <br />

ˆ = ˆ 0 + ˆ 0 (12.21)<br />

X <br />

=1<br />

∇ 2 + (r R) (12.22)<br />

ˆ 0 = ˆ = − ~2 X <br />

∇ 2 (12.23)<br />

2 <br />

=1<br />

I The zero-order Hamiltonian Ĥ 0 :<br />

• The zero-order Hamiltonian ˆ 0 depends only on the electron kinetic energy operator<br />

and the potential energy and describes the rigid molecule (all R are<br />

fixed!) via the unperturbed SE<br />

ˆ 0 el (r; R) = (0) (R) el (r; R) (12.24)


12.4 Radiationless processes in photoexcited molecules 246<br />

• The zero-order electronic wave functions el<br />

(r; R) describe the electrons and<br />

el<br />

their distribution<br />

¯¯ (r; R)¯¯2 for the case of fixed nuclei in the electronic state<br />

|i withtheelectronicenergy el<br />

.<br />

• The el<br />

(r; R) depend on R only as a parameter, not as a variable (we do not<br />

differentiate or integrate with respect to the in the zero-order SE).<br />

• The el<br />

(r; R) form a complete orthonormal system.<br />

• The eigenvalues (0) (R) of the electronic SE are what we have come to know as<br />

the molecular potential energy functions (or hypersurfaces) (R) of the molecule<br />

with all nuclei at rest.<br />

I The perturbation Ĥ 0 :<br />

• The perturbation ˆ 0 is the nuclear kinetic energy operator ˆ .<br />

c) Solution of the complete SE<br />

• To solve the complete SE<br />

³<br />

ˆ 0 + ˆ 0 − ´<br />

(r R) =0 (12.25)<br />

we use the ansatz<br />

(r R) = X (R) · el (r; R) (12.26)<br />

with the expansion coefficients = (R) depending only on R but not on r.<br />

• Inserting the ansatz 12.26 into the complete SE,<br />

³<br />

ˆ 0 + ˆ 0 − ´ X<br />

(R) el<br />

(r; R) =0 (12.27)<br />

and doing some algebraic manipulation, we arrive at a<br />

• system of coupled equations for the electronic wavefunctions el (r; R) for the<br />

electronic state el<br />

(r; R) and the nuclear wavefunctions (R):<br />

and<br />

ˆ 0 el<br />

(r; R) = <br />

(0)<br />

(R) el<br />

(r; R) (12.28)<br />

ˆ 0 (R)+ X (R) (R) = ¡ − (0)<br />

(R) ¢ (R) (12.29)<br />

with the coupling coefficients


12.4 Radiationless processes in photoexcited molecules 247<br />

(R) =<br />

Z<br />

∗ (r; R) ˆ 0 (r; R) r<br />

r<br />

−~ 2 Z<br />

r<br />

∗ (r; R) X <br />

1<br />

<br />

<br />

<br />

(r; R) r<br />

<br />

<br />

(12.30)<br />

• We interprete these Eqs. as follows:<br />

(1) <br />

(0) (R) can be interpreted as the zero-order potential functions of the th<br />

zero-order electronic state that is given by the solution of the electronic SE<br />

(Eq. 12.24) for fixed nuclei R.<br />

(2) Without the (R), Eq. 12.29 describes the kinetic energy of the nuclei<br />

in the potential (0) (R). 52 Eq. 12.29 is thus simply the vibrational SE for<br />

the electronic potential <br />

(0) (R).<br />

(3) The coefficients (R) defined by Eq. 12.30 are coupling terms that connect<br />

(i.e., mix) the electronic states and . They describe how the<br />

different electronic states and are coupled by the nuclear motion (the<br />

two terms on the RHS resulting from the nuclear kinetic energy operator<br />

ˆ ;seeEq.12.30).<br />

d) Born-Oppenheimer approximation<br />

The Born-Oppenheimer (BO) approximation completely neglects the coupling coefficients<br />

. The complete SE is therefore reduced to two uncoupled equations, which<br />

we denote as electronic SE<br />

ˆ 0 el<br />

(r) = <br />

(0)<br />

(R) el<br />

(r) (12.31)<br />

and nuclear (vibrational) SE<br />

h i<br />

ˆ + <br />

(0) (R) = (R) (12.32)<br />

which describe, respectively, the electronic wavefunction el<br />

for fixed nuclear coordinates<br />

R in the electronic state el<br />

and the set of nuclear wavefunctions (R) for the energy<br />

state of the nuclei in the electronic state el<br />

.<br />

e) Breakdown of the Born-Oppenheimer approximnation<br />

When two (or more) electronic states el<br />

(R), el<br />

(R) become nearly isoenergetic or,<br />

at a crossing, even degenerate, the coupling matrix element (R) may no longer be<br />

neglected, because through the (R), thenuclear motion leads to a mixing between<br />

the different electronic BO-states.<br />

52 We used to denote these potential energy functions (≡ potential energy hypersurfaces) as (R).


12.4 Radiationless processes in photoexcited molecules 248<br />

I<br />

This breakdown of the BO approximnation leads to non-adiabatic coupling of<br />

electronic states:<br />

• For a diatomic molecule, the coupling leads to an avoided crossing of the two<br />

electronic potential curves.<br />

• For polyatomic molecules, the coupling can lead to a conical intersection (CI)<br />

between the two electronic potential energy hypersurfaces.<br />

As will be outlined in the following, the existence of conical intersections between different<br />

excited electronic potential energy hypersurfaces and between an excited and<br />

the ground electronic potential energy hypersurface has dramatic consequences for the<br />

dynamics of photochemical reactions.<br />

The recognition since about the years 2000 - 2005 that conical intersections in polyatomic<br />

molecules are the rule rather than the exception has indeed revolutionized our<br />

understanding of photochemical reactions.<br />

12.4.2 Avoided crossings of two potential curves of a diatomic molecule<br />

I<br />

I<br />

Illustration of an avoided crossing using two diabatic model potential functions<br />

+ coupling term:<br />

1 () = 1 [1 − exp [− 1 ( − 1 )]] 2 (12.33)<br />

2 () = 2 + exp [− 2 ( − 2 )] (12.34)<br />

12 () = 1 [1 + tanh [− 2 ( − 12 )]] (12.35)<br />

12 () is just a convenient model function with the physically correct behavior<br />

12 () → 0 for →∞.<br />

Figure 12.3: Diabatic model potential curves for a diatomic molecule.<br />

Avoided Crossing of Two Potentials of a Diatomic Molecule<br />

Diabatic model potential curves: V1<br />

( R) 0 <br />

H <br />

<br />

0 V2<br />

( R)<br />

<br />

1exp ( ) <br />

2<br />

V R D <br />

R R<br />

1 1 1 1<br />

<br />

exp ( ) <br />

V R D R R<br />

2 2 2 2<br />

<strong>Physical</strong> <strong>Chemistry</strong> III: <strong>Chemical</strong> <strong>Kinetics</strong> • © F. Temps, IPC Kiel 254


12.4 Radiationless processes in photoexcited molecules 249<br />

I<br />

Figure 12.4: Avoided crossing of two potential curves of a diatomic molecule.<br />

Avoided Crossing of Two Potentials of a Diatomic Molecule<br />

Adiabatic potential curves: V1<br />

( R) W( R)<br />

<br />

H <br />

<br />

W( R) V2( R)<br />

<br />

V<br />

<br />

<br />

V1<br />

( R) W( R)<br />

<br />

R<br />

Eigenvalues <br />

WR ( ) V2( R)<br />

<br />

<strong>Physical</strong> <strong>Chemistry</strong> III: <strong>Chemical</strong> <strong>Kinetics</strong> • © F. Temps, IPC Kiel 255<br />

I<br />

Figure 12.5: Avoided crossing of two potential curves of a diatomic molecule. At the<br />

avoided crossing the characters of the adiabatic potential curves switch.<br />

Avoided Crossing of Two Potentials of a Diatomic Molecule<br />

V E W V V V V<br />

1 2 1 2 1 2<br />

0 E<br />

W<br />

with: ; <br />

W V2<br />

E<br />

2 2<br />

<br />

2<br />

V ( R) ( R) ( R) W( R)<br />

"Avoided Crossing" with<br />

Splitting of 2 |W( R x ) |<br />

<strong>Physical</strong> <strong>Chemistry</strong> III: <strong>Chemical</strong> <strong>Kinetics</strong> • © F. Temps, IPC Kiel 256


12.4 Radiationless processes in photoexcited molecules 250<br />

I<br />

Figure 12.6: Conical intersection of potential energy surfaces of a polyatomic molecule.<br />

Conical Intersections of Potential Surfaces in Polyatomic Molecules<br />

V E W V V V V<br />

1 2 1 2 1 2<br />

0 E<br />

W<br />

with: ; <br />

W V2<br />

E<br />

2 2<br />

V<br />

S 2<br />

Conical Intersections<br />

Diatomic Molecules: s = 1<br />

<br />

2<br />

V ( R) ( R) ( R) W( R)<br />

S 1<br />

S 0<br />

Q 1<br />

Q 2<br />

Polyatomic Molecules: s = 3N-6<br />

V ( Q , Q ) ( Q , Q )<br />

<br />

1 2 1 2<br />

<br />

( Q , Q ) W( Q , Q ) <br />

<br />

1 2 1 2<br />

2<br />

<strong>Physical</strong> <strong>Chemistry</strong> III: <strong>Chemical</strong> <strong>Kinetics</strong> • © F. Temps, IPC Kiel 257<br />

12.4.3 Quantum mechanical treatment of a coupled 2×2 system<br />

We can easily verify the “avoided crossing” of two coupled potential energy curves as<br />

follows:<br />

I<br />

Schrödinger equation:<br />

( − ) |i =0 (12.36)<br />

I<br />

Hamilton-Operator:<br />

= (0) + (1) (12.37)<br />

I<br />

Ansatz:<br />

|i = 1 | 1 i + 2 | 2 i (12.38)<br />

with orthonormal basis vectors that are eigenvectors of (0) , i.e.,<br />

½<br />

­ ® 1<br />

| = =<br />

0<br />

for = <br />

for 6= <br />

(12.39)<br />

and the zero-order energies<br />

(0) | 1 i = (0)<br />

1 | 1 i (12.40)<br />

(0) | 2 i = (0)<br />

2 | 2 i (12.41)


12.4 Radiationless processes in photoexcited molecules 251<br />

I Solution of the Schrödinger equation: Insertion of the ansatz (Eq. 12.38) into the<br />

SE and multiplication from the left by either h 1 | or by h 2 | gives the two equations:<br />

1 h 1 | ( − ) | 1 i + 2 h 1 | ( − ) | 2 i =0 (12.42)<br />

1 h 2 | ( − ) | 1 i + 2 h 2 | ( − ) | 2 i =0 (12.43)<br />

This is a system of coupled linear equations (“secular equations”):<br />

with the “matrix elements” (integrals over all r resp. R):<br />

Specifically, we have<br />

1 ( 11 − )+ 2 12 =0 (12.44)<br />

1 21 + 2 ( 22 − ) =0 (12.45)<br />

= h | | i (12.46)<br />

= h | (0) + (1) | i (12.47)<br />

= h | (0) | i + h | (1) | i (12.48)<br />

11 = h 1 | (0) | 1 i = (0)<br />

1 (12.49)<br />

22 = h 2 | (0) | 2 i = (0)<br />

2 (12.50)<br />

and<br />

12 = h 1 | (1) | 2 i = h 2 | (1) | 1 i = 21 (12.51)<br />

Therefore, we obtain the secular equations as follows:<br />

I<br />

Secular equations:<br />

³ ´<br />

1 (0)<br />

1 − + 2 12 =0 (12.52)<br />

³ ´<br />

1 21 + 2 (0)<br />

2 − =0 (12.53)<br />

I Secular determinant: A non-trivial solution for the secular equations requires that<br />

(0)<br />

1 − 12<br />

¯ 21 (0)<br />

=0 (12.54)<br />

2 − ¯<br />

I<br />

Eigenvalues:<br />

0=<br />

³ ´³<br />

(0)<br />

1 − <br />

= (0)<br />

1 (0)<br />

2 − <br />

´<br />

(0)<br />

2 − − | 12 | 2 (12.55)<br />

³ ´<br />

(0)<br />

1 + (0)<br />

2 + 2 − | 12 | 2 (12.56)<br />

with<br />

because (1) is Hermitian.<br />

| 12 | 2 = 12 21 (12.57)


12.4 Radiationless processes in photoexcited molecules 252<br />

With some rewriting 53 ,weobtain<br />

± = 1 ³ ´<br />

(0)<br />

1 + (0)<br />

2 ± 1 r ³<br />

(0)<br />

1 + (0)<br />

´2<br />

(0)<br />

2 − 4 1 (0)<br />

2 +4| 12 | 2 (12.58)<br />

2<br />

2<br />

y<br />

± = 1 2<br />

³ ´<br />

(0)<br />

1 + (0)<br />

2 ± 1 r ³<br />

(0)<br />

1 − (0)<br />

´2<br />

2 +4|12 | 2 (12.59)<br />

2<br />

I Shorthand form: Using the abbreviations<br />

± (R) = ± (12.60)<br />

1 (R) = (0)<br />

1 (12.61)<br />

2 (R) = (0)<br />

2 (12.62)<br />

2 (R) =| 12 | 2 (12.63)<br />

Σ (R) = 1 2 ( 1 (R)+ 2 (R)) (12.64)<br />

∆ (R) = 1 2 ( 1 (R) − 2 (R)) (12.65)<br />

we can rewrite the solution for the eigenvalues + and − in a compact form as<br />

± (R) =Σ (R) ±<br />

q<br />

(∆ (R)) 2 +( (R)) 2 (12.66)<br />

12.4.4 Radiationless processes in polyatomic molecules: Conical intersections<br />

I “Traditional” view on the cis-trans isomerization of C 2 H 4 :<br />

• Reaction coordinate can be approximated by torsion angle (0 ≤ ≤ 180 ◦ ).<br />

• Reaction proceeds via a perpendicular transition state ( =90 ◦ ), in which the C<br />

atoms are connected by a single bond.<br />

• Correlation diagram shows that the orbital of the trans-isomer transforms into<br />

the ∗ orbital of the cis-isomer and vice versa.<br />

• Crossing of the two diabatic states corresponds to an avoided crossing of the<br />

adiabatic states.<br />

53<br />

³<br />

(0)<br />

1 + (0)<br />

´2 ³<br />

(0)<br />

2 − 4 1 (0) 2 =<br />

(0)<br />

1<br />

´2<br />

+<br />

³<br />

³<br />

=<br />

(0)<br />

2<br />

(0)<br />

1<br />

´2<br />

+2<br />

(0)<br />

´2<br />

+<br />

³<br />

1 (0) 2 − 4 (0)<br />

1 (0) 2<br />

(0)<br />

2<br />

´2<br />

− 2<br />

(0)<br />

1 (0) 2 =<br />

³<br />

(0)<br />

1 − (0)<br />

´2<br />

2


12.4 Radiationless processes in photoexcited molecules 253<br />

• The reaction is thus thermally forbidden because of the high energy barrier at the<br />

avoided crossing between the trans and cis isomers on the left and right.<br />

• The reaction is photochemically allowed because of the barrierless adiabatic correlation<br />

between the ∗ states on the left and right.<br />

I<br />

Figure 12.7: Cis-trans isomerization of C 2 H 4 : Transformation of the and ∗ orbitals.<br />

Photochemical cis-trans isomerization of C 2 H 4<br />

Traditional picture: Diabatic potentials<br />

<br />

<br />

diabatic crossing<br />

<br />

Topics in <strong>Chemical</strong> <strong>Kinetics</strong> and Dynamics • © F. Temps, IPC Kiel 35<br />

I<br />

Figure 12.8: Cis-trans isomerization of C 2 H 4 : HMO picture with an avoided crossing.<br />

Photochemical cis-trans isomerization of C 2 H 4<br />

Traditional picture: Avoided crossing<br />

<br />

<br />

avoided crossing<br />

<br />

Topics in <strong>Chemical</strong> <strong>Kinetics</strong> and Dynamics • © F. Temps, IPC Kiel 36


12.4 Radiationless processes in photoexcited molecules 254<br />

I<br />

Figure 12.9: Internal coordinates relevant for the cis-trans isomerization of C 2 H 4 .The<br />

out-of-plane bending angle plays a role as coupling mode leading to a CI; the angles<br />

and play a role in the isomerization to H 3 CCH (also by a CI).<br />

Photochemical cis-trans isomerization of C 2 H 4<br />

Important internal coordinates:<br />

Torsion<br />

Out-of-plane bend<br />

Barbatti et al., 2004<br />

<strong>Physical</strong> <strong>Chemistry</strong> III: <strong>Chemical</strong> <strong>Kinetics</strong> • © F. Temps, IPC Kiel 258<br />

I Figure 12.10: Diabatic model PES’s for C 2 H 4 .<br />

Photochemical cis-trans isomerization of C 2 H 4<br />

Diabatic PES:<br />

<br />

(s1)-dimensional<br />

crossing seem<br />

<br />

<br />

<br />

<br />

<br />

<strong>Physical</strong> <strong>Chemistry</strong> III: <strong>Chemical</strong> <strong>Kinetics</strong> • © F. Temps, IPC Kiel 259


12.4 Radiationless processes in photoexcited molecules 255<br />

I Figure 12.11: Adiabatic model PES’s for C 2 H 4 with the conical intersection at =90 ◦ .<br />

Photochemical cis-trans isomerization of C 2 H 4<br />

Adiabatic PES:<br />

<br />

<br />

conical intersection<br />

(s-2) dimensional<br />

<br />

<br />

<br />

<br />

<strong>Physical</strong> <strong>Chemistry</strong> III: <strong>Chemical</strong> <strong>Kinetics</strong> • © F. Temps, IPC Kiel 260<br />

I<br />

Figure 12.12: Wave packet motion in a complex photo-induced reaction.<br />

Dynamics of Photo-Induced Reactions: Photochemical Funnels<br />

Conical Intersection =<br />

"Photochemical Funnel"<br />

Energy<br />

Q b<br />

S 0<br />

S 1<br />

S 0<br />

Reaction Coordinate<br />

<strong>Physical</strong> <strong>Chemistry</strong> III: <strong>Chemical</strong> <strong>Kinetics</strong> • © F. Temps, IPC Kiel 262


12.4 Radiationless processes in photoexcited molecules 256<br />

13. Atmospheric chemistry*<br />

(skipped for lack of time in the semester)


13 Atmospheric chemistry* 257<br />

14. Combustion chemistry*<br />

(skipped for lack of time in the semester)


14 Combustion chemistry* 258<br />

15. Astrochemistry*<br />

(skipped for lack of time in the semester)


15 Astrochemistry* 259<br />

16. Energy transfer processes*<br />

(See lecture in M.Sc. course.)


16 Energy transfer processes* 260<br />

17. Reaction dynamics*<br />

(See lecture in M.Sc. course.)


17 Reaction dynamics* 261<br />

18. Electrochemical kinetics*<br />

(skipped for lack of time in the semester)


Appendix B 262<br />

Appendix A: Useful physicochemical constants<br />

I Table A.1: List of useful physical constants (Cohen 1986).<br />

physical constant<br />

symbol value<br />

Loschmidt (or Avogadro) number =60221367 × 10 23 mol −1<br />

gas constant =8314511 J mol −1 K −1<br />

Boltzmann constant =1380658 × 10 −23 JK −1<br />

electron charge =160217733 × 10 −19 C<br />

Faraday constant = 96485309 C mol −1<br />

speed of light =299792458 × 10 8 ms −1<br />

Planck’s constant =66260755 × 10 −34 Js<br />

~ ~ =10547266 × 10 −34 Js<br />

electron mass =91093897 × 10 −31 kg<br />

proton mass =16726231 × 10 −27 kg<br />

neutron mass =16749286 × 10 −27 kg<br />

atomic mass unit =16605402 × 10 −27 kg<br />

Rydberg constant e e =10973731534 × 10 7 m −1<br />

Bohr radius 0 0 =529177249 × 10 −11 m<br />

electric field constant 0 0 =1( 0 2 )<br />

0 =8854187817 × 10 −12 Fm −1<br />

magnetic field constant 0 0 =4 × 10 −7 NA −2<br />

0 =125664 × 10 −6 NA −2<br />

I<br />

References:<br />

Cohen 1986 E. R. Cohen and B. N. Taylor, The 1986 Adjustment of the Fundamental<br />

Constants, CODATA Bull. 63, 1 (1986). An updated list is contained in every<br />

August issue of Physics Today.


Appendix B 263<br />

Appendix B: The Marquardt-Levenberg non-linear least-squares<br />

fitting algorithm<br />

With the omnipresence of the computer, the Marquardt-Levenberg algorithm (Bevington<br />

1992, Press 1992) has become an extremely important data analysis tool. It is generally<br />

themethodofchoiceincaseswhereitisnotpossibletomakesimplelinearplotsto<br />

determine rate constants.<br />

I Problem: Experimentally, we measure the values of some quantity at the points ,<br />

, . Supposewetake measurements.Nowsupposewewouldliketodescribeour<br />

data points by some model. Towards these ends, we take a physically reasonable<br />

model function<br />

= ( ;( 1 ))<br />

(B.1)<br />

which we would like to fit to our data in such a way that () describes the data<br />

points in “the best possible way”. The model function depends directly on the<br />

variables , and it also depends parametrically on nonlinear parameters .In<br />

kinetics, for example, is usually some concentration, e.g. , which depends on time<br />

i.e., = (). The parameters may be the rate constants (or decay times )<br />

and the initial concentration 0 . Our job is to adjust and optimize these parameters<br />

by somehow “fitting” them. The method of choice is, of course, “least-squares fitting”:<br />

We start by defining the so-called figure-of-merit or goodness-of-fit function 54<br />

2 =<br />

"<br />

#<br />

X [ − ( )] 2<br />

=1<br />

<br />

2<br />

(B.2)<br />

Here, the and the independent variables , are the measured quantities, the<br />

are the experimental uncertainties of the ,and ( ) is the calculated value<br />

of the model fit function at the point { } (the parameters 1 are<br />

omitted here in ). What we have to do is to adjust the parameters in the model<br />

function such that 2 reaches a minimum, i.e. we have to search the -dimensional<br />

parameter space for the (global) minimum of 2 .<br />

I Solution: The Marquardt-Levenberg algorithm is a method to determine this minimum<br />

of 2 as function of the fit parameters in the model fit function.<br />

The minimum of 2 as function of the fit parameters a is defined by the conditions<br />

"<br />

#<br />

2<br />

= X [ − ( )] 2<br />

=0 (B.3)<br />

<br />

<br />

2<br />

=1<br />

<br />

X<br />

∙ ¸<br />

[ − ( )] ( )<br />

= −2<br />

(B.4)<br />

<br />

2 <br />

=1<br />

Thus, taking the partial derivatives of 2 with respect to each of the parameters<br />

,weobtainasetof coupled equations in the unknown parameters . These<br />

equations are, in general, nonlinear in the .<br />

54 Deutsch: Fehlerquadratsumme.


Appendix B 264<br />

The Marquardt-Levenberg algorithm gives a recipy for finding the minimum of 2 using<br />

a combination of the method of steepest descent (following the gradient of the hypersurface<br />

defined by 2 as a function of the parameters 1 and, near the minimum<br />

of 2 , a parabolic Taylor series expansion of the fitting function around the minimum).<br />

At the end of a fit, we shall usually report the (± 2 standard deviations), the<br />

standard deviation of the ’s, and the so-called reduced- 2 ,whichis 2 divided by the<br />

number of degrees of freedom, = − , i.e.,<br />

"<br />

#<br />

2 = 1 X [ − ( )] 2<br />

(B.5)<br />

− <br />

<br />

2 <br />

=1<br />

I Example 1: Exponential decay. Suppose we measure the time dependent concentration<br />

() of a molecule in some quantum state, which is populated at some time 0<br />

by a laser pulse and then decays monoexponentially. As a model fit function, we use<br />

the expression<br />

() = 1 exp [− 2 ( − 3 )] + 4 + 5 <br />

(B.6)<br />

The parameters we are interested in are 2 (the decay rate coefficient) and 1 (the<br />

initial amplitude or initial concentration). The parameter 3 just describes the trigger<br />

delay time (= 0 ). We have also added the parameters 4 and 5 to describe a constant<br />

background term (due to some offset of our experimental signal) and a linear drift in<br />

this background term (perhaps due to some experimental instability, such as a gradual<br />

temperature increase in the lab). Our figure of merit is then<br />

" #<br />

X<br />

2 [ − ()] 2<br />

=<br />

(B.7)<br />

=1<br />

The (in general nonlinear) conditions for the minimum of 2 , from which we determine<br />

the parameters ,are<br />

" #<br />

2<br />

= X [ − ()] 2<br />

X<br />

∙ ¸<br />

[ − ()] ()<br />

= −2<br />

=0<br />

1 1 <br />

2<br />

=1 <br />

2<br />

=1 1<br />

(B.8)<br />

2<br />

2<br />

= (B.9)<br />

<br />

(B.10)<br />

Figure B.1 below shows a simple two-parameter fit (parameters 1 and 1) tosucha<br />

signal.<br />

<br />

2


Appendix B 265<br />

I<br />

Figure B.1: Exponential decay curve of a particular vibration-rotation state of the<br />

CH 3 O radical resulting from the unimolecular dissociation reaction of the radical according<br />

to CH 3 O → H+H 2 CO (Dertinger 1995). The small box is the output box<br />

from a fit usingtheORIGINprogram.<br />

I Example 2: Multiexponential decay. In femtosecond spectroscopy, we often observe<br />

ultrafast multiexponential decays of laser-excited molecules. The laser prepares<br />

an excited wavepacket, which usually does not decay single-exponentially. Further, we<br />

need to take into account the final duration of the pump laser pulse (by deconvolution,<br />

or by forward convolution).<br />

• Model function to be fitted to measured data:<br />

() = X exp (− )<br />

(B.11)<br />

with adjustable parameters and .<br />

• Instrument response function (IRF): Often represented by a Gaussian centered at<br />

time 0 Ã !<br />

1<br />

() = √ exp − ( − 0) 2<br />

(B.12)<br />

IRF 2 2 2 IRF<br />

with width parameter IRF related to the full width at half maximum (FWHM)<br />

of the IRF by<br />

FWHM = √ 8ln2≈ 2355 IRF<br />

(B.13)


Appendix C 266<br />

• Convolution of the molecular intensity () and () gives the signal function<br />

() =<br />

Z<br />

+∞<br />

−∞<br />

( 0 ) ( − 0 ) 0 =<br />

where ⊗ denotes the convolution.<br />

Z<br />

+∞<br />

−∞<br />

( − 0 ) ( 0 ) 0 = () ⊗ ()<br />

(B.14)<br />

• Resulting model function to be fitted to the data:<br />

() = 1 X<br />

∙ 1 2 IRF<br />

exp − ( − ¸ ∙ µ ¸<br />

0)<br />

( − 0 ) − 2 IRF<br />

1+erf √ + <br />

2<br />

2 2 <br />

<br />

2IRF <br />

(B.15)<br />

where erf () is the error function and is a simple constant background term (can<br />

be replaced by background + drift + ).<br />

I<br />

Figure B.2: Excited-state relaxation dynamics of the adenine dinucleotide after UV<br />

photoexcitation.<br />

I<br />

References:<br />

Bevington 1992 P. R. Bevington, D. K. Robinson, Data Reduction and Error Analysis<br />

for the <strong>Physical</strong> Sciences, McGraw-Hill, Boston, 1992.<br />

Dertinger 1995 S. Dertinger, A. Geers, J. Kappert, F. Temps, J. W. Wiebrecht,<br />

Rotation-Vibration State Resolved Unimolecular Dynamics of Highly Vibrationally<br />

Excited CH 3 O( 2 ): III. State Specific Dissociation Rates from Spectroscopic Line<br />

Profiles and Time Resolved Measurements, Faraday Discuss. Roy. Soc. 102, 31<br />

(1995).<br />

Press 1992 W. H. Press, S. A. Teukolsky, W. T. Vetterling, B. P. Flannery, Numerical<br />

Recipes in Fortran, Cambridge University Press, Cambridge, 1992. Versions are<br />

also available for C and Pascal.


Appendix C 267<br />

Appendix C: Solution of inhomogeneous differential equations<br />

C.1 General method<br />

An inhomogeneous DE is a DE of the type<br />

0 + () × = ()<br />

(C.1)<br />

In order to solve Eq. C.1, we first find a solution of the homogeneous DE<br />

0 + () × =0<br />

(C.2)<br />

and then determine a particular solution of the inhomogeneous DE:<br />

• Solution of the homogeneous DE by separation of variables:<br />

0 + () × =0<br />

0 = − () × <br />

(C.3)<br />

(C.4)<br />

y<br />

ln <br />

<br />

= − () (C.5)<br />

= × − ()<br />

(C.6)<br />

• Determination of a particular solution of the inhomogeneoues DE by variation of<br />

constant:<br />

→ ()<br />

(C.7)<br />

0 () =<br />

(C.8)<br />

Insertion of () and 0 () into Eq. C.1 and integration gives<br />

= <br />

(C.9)<br />

The general solution of Eq. C.1 is then given by<br />

= + <br />

(C.10)


Appendix C 268<br />

C.2 Application to two consecutive first-order reactions<br />

Consider the reaction system<br />

A 1<br />

−→ B<br />

B 2<br />

−→ C<br />

(C.11)<br />

(C.12)<br />

which is described by the following rate equations<br />

[A]<br />

<br />

[B]<br />

<br />

[C]<br />

<br />

= − 1 [A] (C.13)<br />

=+ 1 [A] − 2 [B]<br />

(C.14)<br />

=+ 2 [B]<br />

(C.15)<br />

The solution for Eq. C.13 is<br />

[A] = [A] 0<br />

− 1 <br />

(C.16)<br />

Thus we have to find the solution of the inhomogeneous DE for [B] (Eq. C.14)<br />

[B]<br />

<br />

+ 2 [B] = 1 [A] 0<br />

− 1 <br />

(C.17)<br />

• Solution of the homogeneous DE by separation of variables:<br />

y<br />

[B]<br />

<br />

= − 2 [B] (C.18)<br />

[B] <br />

= × − 2 <br />

(C.19)<br />

• Determination of a particular solution by variation of constant:<br />

y<br />

[B] <br />

<br />

= ()<br />

[B] <br />

= () × − 2 <br />

= ()<br />

<br />

× − 2 − () × 2 − 2 <br />

(C.20)<br />

(C.21)<br />

(C.22)<br />

Insertion of [B] <br />

and [B] <br />

<br />

into Eq. C.17 gives<br />

()<br />

<br />

which simplifies to<br />

× − 2 − () × 2 − 2 + () × 2 − 2 = 1 [A] 0<br />

− 1 <br />

()<br />

<br />

= 1 [A] 0<br />

( 2− 1 )<br />

This DE is easily integrated to obtain ():<br />

(C.23)<br />

(C.24)


Appendix C 269<br />

<strong>—</strong> If 1 6= 2 we obtain<br />

and thus<br />

() = 1 [A] 0<br />

2 − 1<br />

( 2− 1 )<br />

(C.25)<br />

[B] <br />

= () − 2 <br />

= 1 [A] 0<br />

2 − 1<br />

( 2− 1 ) − 2 <br />

= 1 [A] 0<br />

2 − 1<br />

− 1 <br />

(C.26)<br />

(C.27)<br />

(C.28)<br />

<strong>—</strong> If 1 = 2 we obtain<br />

()<br />

<br />

= 1 [A] 0<br />

( 2− 1 )<br />

(C.29)<br />

= 1 [A] 0<br />

(C.30)<br />

y<br />

y<br />

() = 1 [A] 0<br />

<br />

(C.31)<br />

[B] <br />

= () − 2 <br />

= 1 [A] 0<br />

− 2 <br />

(C.32)<br />

(C.33)<br />

• General solution:<br />

y<br />

Initial value condition at =0:<br />

[B] = [B] <br />

+[B] <br />

(C.34)<br />

⎧<br />

⎪⎨ × − 2 + 1 [A] 0<br />

− 1 <br />

if 1 6= 2<br />

[B] =<br />

2 − 1<br />

(C.35)<br />

⎪⎩<br />

× − 2 + 1 [A] 0<br />

− 2 <br />

if 1 = 2<br />

[B ( =0)]=0 (C.36)<br />

y<br />

⎧<br />

⎪⎨ − 1 [A] 0<br />

if 1 6= 2<br />

=<br />

2 − 1<br />

(C.37)<br />

⎪⎩<br />

0 if 1 = 2<br />

Final solutions for [B ()]:<br />

⎧<br />

1 [A] ⎪⎨ 0<br />

¡<br />

<br />

− 1 − ¢ − 2 <br />

if 1 6= 2<br />

[B ()] =<br />

2 − 1<br />

(C.38)<br />

⎪⎩<br />

1 [A] 0<br />

− 2 <br />

if 1 = 2


Appendix C 270<br />

C.3 Application to parallel and consecutive first-order reactions<br />

Consider the reaction system<br />

A 1<br />

−→ P<br />

A 1<br />

−→ B 2<br />

−→ P<br />

(C.39)<br />

(C.40)<br />

which is described by the following rate equations<br />

[A]<br />

<br />

[B]<br />

<br />

[P]<br />

<br />

= − ( 1 + 1 )[A]=− 1 [A] (C.41)<br />

=+ 1 [A] − 2 [B]<br />

(C.42)<br />

=+ 1 [A] + 2 [B]<br />

(C.43)<br />

with 1 =( 1 + 1 ). 55<br />

The solution for Eq. C.41 is<br />

[A] = [A] 0<br />

−( 1+ 1 ) =[A] 0<br />

− 1 <br />

(C.44)<br />

Thus we have to find the solution of the inhomogeneous DE for [B] (Eq. C.42)<br />

[B]<br />

+ 2 [B] = + 1 [A]<br />

<br />

0<br />

− 1 <br />

using the above standard procedure:<br />

(C.45)<br />

• Solution of the homogeneous DE by separation of variables:<br />

y<br />

[B]<br />

<br />

= − 2 [B] (C.46)<br />

[B] <br />

= × − 2 <br />

(C.47)<br />

• Determination of a particular solution by variation of constant:<br />

y<br />

[B] <br />

<br />

= ()<br />

[B] <br />

= () × − 2 <br />

= ()<br />

<br />

× − 2 − () × 2 − 2 <br />

(C.48)<br />

(C.49)<br />

(C.50)<br />

55 An example is the radiationless deactivation of a ∗ electronically excited molecule A directly to<br />

the ground state P or via an intermediatate optically dark ∗ state B, which decays to the ground<br />

statemoreslowly.


Appendix C 271<br />

Insertion of [B] <br />

and [B] <br />

<br />

into Eq. C.45 gives<br />

()<br />

<br />

× − 2 − () 2 − 2 + 2 () − 2 = 1 [A] 0<br />

− 1 <br />

(C.51)<br />

The terms with ± () 2 − 2 cancel, so that<br />

()<br />

× − 2 = 1 [A]<br />

<br />

0<br />

− 1 <br />

which we rewrite in order to solve () as<br />

()<br />

<br />

= 1 [A] 0<br />

− 1 × + 2 <br />

or<br />

()<br />

= 1 [A]<br />

<br />

0<br />

( 2− 1 ) <br />

This DE is easily integrated to obtain ():<br />

(C.52)<br />

(C.53)<br />

(C.54)<br />

<strong>—</strong> If 2 6= 1 =( 1 + 1 ) we obtain<br />

() = 1 [A] 0<br />

2 − 1<br />

( 2− 1 )<br />

(C.55)<br />

and thus<br />

[B] <br />

= () − 2 <br />

= 1 [A] 0<br />

2 − 1<br />

( 2− 1 ) − 2 <br />

= 1 [A] 0<br />

2 − 1<br />

− 1 <br />

(C.56)<br />

(C.57)<br />

(C.58)<br />

<strong>—</strong> The case 2 = 1 does not interest us here, as the ∗ state B should be a<br />

longer-lived one.<br />

• General solution for 2 6= 1 :<br />

y<br />

Initial value condition at =0:<br />

[B] = [B] <br />

+[B] <br />

[B] = × − 2 + 1 [A] 0<br />

2 − 1<br />

− 1 <br />

(C.59)<br />

(C.60)<br />

[B ( =0)]=0 (C.61)<br />

y<br />

= − 1 [A] 0<br />

2 − 1<br />

(C.62)


Appendix D 272<br />

Final solutions for [B ()]:<br />

i.e.<br />

[B ()] = − 1 [A] 0<br />

2 − 1<br />

× − 2 + 1 [A] 0<br />

2 − 1<br />

− 1 <br />

= 1 [A] 0<br />

2 − 1<br />

− 1 − 1 [A] 0<br />

2 − 1<br />

× − 2 <br />

[B ()] = 1 [A] 0<br />

1 − 2<br />

× ¡ − 2 − − 1 ¢<br />

(C.63)<br />

(C.64)<br />

(C.65)<br />

• In the time-resolved fluorescence measurements, we observe<br />

() ∝ [A ()] + [B ()]<br />

(C.66)<br />

with some unknown factor accounting for the (lower) transition probability of<br />

state B, i.e., the fluorescence-time profile () is<br />

() =[A] 0<br />

− 1 + 1 [A] 0<br />

1 − 2<br />

× ¡ − 2 − − 1 ¢<br />

=[A] 0<br />

− 1 + 1 [A] 0<br />

1 − 2<br />

× − 2 − 1 [A] 0<br />

1 − 2<br />

× − 1 <br />

= µ<br />

1 [A] 0<br />

× − 2 + [A]<br />

1 − 0<br />

− <br />

1 [A] 0<br />

× − 1 <br />

2 1 − 2<br />

= 2 × − 2 + 1 × − 1 <br />

(C.67)<br />

(C.68)<br />

(C.69)<br />

(C.70)<br />

This is the same result as would be obtained by a fit toasumoftwoexponentials<br />

with the rate constants 1 = 1 + 1 and 2 , i.e., time constants<br />

1 =( 1 + 1 ) −1 and 2 = 2 .


Appendix D 273<br />

Appendix D: Matrix methods<br />

This appendix has been taken from the “Introduction to Molecular Spectroscopy” script,<br />

not all subsections apply to chemical kinetics.<br />

D.1 Definition<br />

I Matrices: A matrix is a rectangular, × dimensional array of numbers :<br />

A =<br />

⎛<br />

⎜<br />

⎝<br />

⎞<br />

11 12 13 1<br />

21 22 23 2<br />

⎟<br />

. . . . ⎠<br />

1 2 3 <br />

(D.1)<br />

The are called the matrix elements.<br />

We will need to consider only × dimensional square matrices, which we can think<br />

of as a collection of column vectors a :<br />

⎛<br />

⎞<br />

11 12 1<br />

<br />

A = ⎜ 21 22 2<br />

⎟<br />

⎝ . . . ⎠<br />

(D.2)<br />

1 2 <br />

I Notation: Matrices will be denoted by bold symbols:<br />

A<br />

(D.3)<br />

I Matrix manipulation by computer programs: Matrix manipulations are carried<br />

out efficiently using computer programs (see, e.g., Press1992) or with symbolic algebra<br />

programs. 56<br />

56 The most commmon symbolic math programs are MathCad, MuPad, Maple, andMathematica.<br />

MathCad and Mathematica are available in the PC lab.


Appendix D 274<br />

D.2 Special matrices and matrix operations<br />

I Identity matrix: I = 1<br />

I = 1 =<br />

⎛<br />

⎜<br />

⎝<br />

10 0<br />

01 0<br />

. . .<br />

00 1<br />

⎞<br />

⎟<br />

⎠<br />

(D.4)<br />

We frequently use the Kronecker :<br />

= <br />

(D.5)<br />

where<br />

½<br />

1 for = <br />

=<br />

0 for 6= <br />

(D.6)<br />

I<br />

Diagonal matrices:<br />

⎛<br />

⎜<br />

⎝<br />

⎞<br />

11 0<br />

0 22 0<br />

⎟<br />

. . . . ⎠<br />

0 0 <br />

(D.7)<br />

I<br />

Block diagonal matrices:<br />

⎛<br />

⎞<br />

11 12 0<br />

<br />

⎜ 21 22 0<br />

⎟<br />

⎝ . . . . ⎠<br />

0 0 <br />

(D.8)<br />

I Complex conjugate (c.c.) of a matrix: A ∗<br />

To obtain the complex conjugate A ∗ of the matrix A, change to −:<br />

( ∗ ) <br />

=( ) ∗ (D.9)<br />

I Transpose of a matrix: <br />

To obtain the transpose A of the matrix A, interchange rows and colums:<br />

¡<br />

<br />

¢ =( ) (D.10)


Appendix D 275<br />

I Hermitian conjugate (“transpose conjugate”) of a matrix: A †<br />

To obtain the Hermitian conjugate (or “transpose conjugate” or “adjoint”) A † of the<br />

matrix A, take the complex conjugate and transpose:<br />

† = ¡ ∗¢ = ¡ ¢ ∗<br />

(D.11)<br />

where ¡<br />

<br />

† ¢ =( ) ∗ (D.12)<br />

I Hermitian matrices: AmatrixA is Hermitian, if<br />

A † =(A ∗ ) = A<br />

(D.13)<br />

i.e., ¡<br />

<br />

† ¢ =( ) ∗ = (D.14)<br />

I Inverse of a matrix: A −1<br />

AmatrixA −1 is the inverse of the original matrix A, if<br />

AA −1 = A −1 A = I<br />

(D.15)<br />

I Orthogonal matrices: AmatrixA is orthogonal, iftheinverseofA equals its transpose:<br />

A −1 = A <br />

(D.16)<br />

I Unitary matrices: AmatrixU is unitary, iftheinverseofU equals its transpose<br />

conjugate:<br />

U −1 = U † =(U ∗ ) <br />

(D.17)<br />

y<br />

U † U = I<br />

(D.18)<br />

I<br />

Trace of a matrix:<br />

tr (A) =<br />

X<br />

=1<br />

<br />

(D.19)<br />

I<br />

Matrix addition:<br />

C = A + B<br />

(D.20)<br />

with<br />

= + <br />

(D.21)


Appendix D 276<br />

I<br />

Multiplication by a scalar:<br />

C = A<br />

(D.22)<br />

with<br />

= <br />

(D.23)<br />

I<br />

Matrix multiplication:<br />

C = A · B = AB<br />

(D.24)<br />

with<br />

=<br />

X<br />

<br />

(D.25)<br />

=1<br />

Example:<br />

µ µ µ µ <br />

12 56 1 × 5+2× 71× 6+2× 8 19 22<br />

=<br />

34 78 3 × 5+4× 73× 6+4× 8 43 50<br />

(D.26)<br />

Notes:<br />

(1) Matrix multiplication is defined only if the number of columns of the first matrix<br />

is identical to the number of rows of the second matrix.<br />

(2)<br />

AB6= BA<br />

(D.27)<br />

I Singular matrices: AmatrixA is called singular, if its determinant vanishes:<br />

det (A) =|A| =0<br />

(D.28)


Appendix D 277<br />

D.3 Determinants<br />

I Determinant of a matrix: A determinant of a matrix is, in general, a number which<br />

is evaluated according to the rule<br />

det (A)=|A| =<br />

X<br />

<br />

=1<br />

(D.29)<br />

The are called the co-factors which are obtained by omitting the ’th row and ’th<br />

column (marked below in red) and multiplying by (−1) + :<br />

11 1 1<br />

. . .<br />

=(−1)<br />

¯¯¯¯¯¯¯¯¯¯¯<br />

+ 1 <br />

. . .<br />

1 <br />

¯¯¯¯¯¯¯¯¯¯¯<br />

(D.30)<br />

I<br />

Evaluation of simple determinants:<br />

¯ <br />

¯ = − <br />

(D.31)<br />

11 12 13<br />

21 22 23 = (D.32)<br />

¯ 31 32 33<br />

¯¯¯¯¯¯<br />

= 11 22 33 + 12 23 31 + 13 21 32<br />

− ( 13 22 31 + 11 23 32 + 12 21 33 )<br />

where we “extended” the original determinant according to the scheme<br />

11 12 13 11 12<br />

21 22 23 21 22<br />

(D.33)<br />

¯ 31 32 33<br />

¯¯¯¯¯¯ 31 32


Appendix D 278<br />

D.4 Coordinate transformations<br />

I Multiplication of a vector by a matrix: x 0 = Ax<br />

The multiplication of an -dimensional column vector x by an × matrix A gives a<br />

new vector x 0 with components corresponding to those of x in a transformed coordinate<br />

system.<br />

I Example: Multiplication of a 2-dim vector<br />

x =<br />

µ<br />

1<br />

2<br />

<br />

(D.34)<br />

by a matrix<br />

gives<br />

µ <br />

cos − sin <br />

A =<br />

sin cos <br />

(D.35)<br />

µ µ <br />

cos − sin 1<br />

Ax =<br />

(D.36)<br />

sin cos 2<br />

µ <br />

1 cos − <br />

=<br />

2 sin <br />

(D.37)<br />

1 sin + 2 cos <br />

= x 0 (D.38)<br />

I<br />

Figure D.1: Rotation of a coordinate system.


Appendix D 279<br />

I<br />

The result is a rotation:<br />

• Intheoriginalcoordinatesystem,thenewvectorx 0 corresponds to the result of<br />

an anti-clockwise rotation of x around .<br />

• Alternatively, we can say that x 0 isgiveninanewcoordinatesystemthatis<br />

obtained by a clockwise rotation of the old coordinate system around .<br />

I Rotation matrices and rotation operators: A is called a rotation matrix. Weshall<br />

also call A a rotation operator.<br />

I<br />

Exercise D.1: Show that Eqs. D.36 - D.38 does describe an anti-clockwise rotation<br />

of the vector x to the new vector x 0 and determine its coordinates using trigonometric<br />

arguments. ¤


Appendix D 280<br />

D.5 Systems of linear equations<br />

I Solutions of linear equations (I): Cramer’s rule. Matrices and determinants can<br />

be generally used for solving systems of linear equations of the type<br />

11 1 + 12 2 + 13 3 + 1 = 1<br />

21 1 + 22 2 + 23 3 + 2 = 2<br />

= <br />

(D.39)<br />

1 1 + 2 2 + 3 3 + <br />

= <br />

These equations can be written in matrix form:<br />

⎛<br />

⎞ ⎛ ⎞ ⎛ ⎞<br />

11 12 1 1 1<br />

<br />

⎜ 21 22 2<br />

<br />

⎟ ⎜ 2<br />

⎟<br />

⎝ . . . ⎠ ⎝ . ⎠ = <br />

⎜ 2<br />

⎟<br />

⎝ . ⎠<br />

1 2 <br />

or<br />

Ax= c<br />

(D.40)<br />

(D.41)<br />

Simple algebraic manipulations (see any math textbook) gives the expressions<br />

|A| = |A |<br />

(D.42)<br />

where |A| is the determinant of the coefficient matrix A and |A | is the respective<br />

determinant of |A| in which the th column is replaced by the 1 2 3 ,e.g.<br />

|A 2 | =<br />

⎛<br />

⎞<br />

11 1 1<br />

<br />

⎜ 21 2 2<br />

⎟<br />

⎝ . . . ⎠<br />

1 <br />

(D.43)<br />

Solutions for the : From Eq. D.42 we obtain the non-trivial solutions for the <br />

according to<br />

= |A |<br />

(D.44)<br />

|A|<br />

under the condition that the A matrix is not singular, i.e., the determinant of coefficients<br />

does not vanish<br />

|A| 6= 0<br />

(D.45)<br />

(The trivial and uninteresting solutions are 1 2 3 =0).<br />

I Solutions of linear equations (II): In the following, we shall only consider the special<br />

linear equations of the type<br />

11 1 + 12 2 + 13 3 + 1 =0<br />

21 1 + 22 2 + 23 3 + 2 =0<br />

= .<br />

1 1 + 2 2 + 3 3 + =0<br />

(D.46)


Appendix D 281<br />

or<br />

Ax=0<br />

(D.47)<br />

The trivial and uninteresting solutions of this equation are 1 2 3 =0.<br />

The interesting solutions require that the matrix of coefficients A is singular, i.e., the<br />

determinant of coefficients has to vanish<br />

|A| =0<br />

(D.48)<br />

This type of equations is encountered in eigenvalue problems.


Appendix D 282<br />

D.6 Eigenvalue equations<br />

Consider a set of linear equations as considered in the previous section which can be<br />

writtenintheform<br />

Ax= x<br />

(D.49)<br />

where<br />

(1) A is an × -dimensional matrix,<br />

(2) is a scalar constant, called an eigenvalue of A ( is any one of the set of <br />

eigenvalues of A), and<br />

(3) x is an -dimensional column vector (which can in general be complex), called<br />

the eigenvector of A belonging to the particular eigenvalue.<br />

Eq. D.49 is called an eigenvalue equation. It has the following property: The multiplication<br />

of x by (and hence that of x by the matrix A) changes the length of x (by<br />

the factor ), but not the direction.<br />

I Eigenvalues and eigenvectors of the molecular Hamiltonian: The determination<br />

of eigenvalues and eigenvectors of the molecular Hamiltonian H, i.e., the “energy<br />

matrix”, or the matrix of the Hamilton operator b , is the most important problem in<br />

spectroscopy (in general, it is the key problem!!).<br />

I Secular equation and eigenvalues λ : Eq. D.49 can be rewritten as<br />

(A − I) x = 0 (D.50)<br />

In order not to obtain the trivial solutions 1 = 2 = 3 = =0, det (A − I) has<br />

to vanish, i.e., det (A − I) =0. This is the so-called characteristic equation or<br />

secular equation:<br />

|A − I| =<br />

¯<br />

The secular equation has roots<br />

which are called the eigenvalues.<br />

11 − 12 1<br />

21 22 − 2<br />

. . .<br />

=0 (D.51)<br />

1 2 − ¯<br />

{ 1 2 } (D.52)


Appendix D 283<br />

I Eigenvectors x : For each eigenvalue , we can write an eigenvalue equation<br />

In matrix notation, this becomes<br />

Ax 1 = 1 x 1 (D.53)<br />

Ax 2 = 2 x 2 (D.54)<br />

. (D.55)<br />

Ax = x (D.56)<br />

with the diagonal eigenvalue matrix<br />

Λ =<br />

AX= X Λ<br />

⎛<br />

⎜<br />

⎝<br />

⎞<br />

1 0 0<br />

0 2 0<br />

⎟<br />

. . . ⎠<br />

0 0 <br />

(D.57)<br />

(D.58)<br />

The different eigenvectors x are usually normalized to unity, i.e.,<br />

|x| = ¡ x † x ¢ 12<br />

=1<br />

(D.59)<br />

I Matrix diagonalization: The matrix of the eigenvectors X can be determined by<br />

multiplying the equation<br />

AX= X Λ<br />

(D.60)<br />

from the left with X −1 ,whichgives<br />

X −1 AX= X −1 X Λ<br />

(D.61)<br />

i.e.,<br />

y<br />

X −1 AX= Λ<br />

(D.62)<br />

The determination of the eigenvector matrix is equivalent to finding a matrix X that<br />

transforms the matrix A into diagonal form (Λ).<br />

I<br />

Example:<br />

A =<br />

µ <br />

2 3<br />

310<br />

(D.63)<br />

Eigenvalue equations:<br />

Ax = x <br />

(D.64)<br />

or in matrix form:<br />

AX= X Λ<br />

(D.65)


Appendix E 284<br />

Secular equation:<br />

det (A − I) =<br />

¯ 2 − 3<br />

3 10− ¯<br />

=(2− )(10− ) − 9=0<br />

(D.66)<br />

(D.67)<br />

y<br />

0=20− 2 − 10 + 2 − 9= 2 − 12 +11 (D.68)<br />

12 =+ 12 2 ± s µ12<br />

2<br />

2<br />

− 11 = 6 ± √ 36 − 11 (D.69)<br />

=6± 5 (D.70)<br />

y<br />

Eigenvalues:<br />

1 =1 (D.71)<br />

2 =11 (D.72)<br />

Eigenvectors: Inserting the eigenvalues one by one into the eigenvalue equation yields:<br />

(1) µ µ µ µ <br />

2 3 11 11 11<br />

= <br />

310 1 =11<br />

21 21 21<br />

y<br />

µ<br />

−3<br />

1<br />

(2) µ µ µ µ <br />

2 3 12 12 12<br />

= <br />

310 2 =1<br />

22 22 22<br />

y µ <br />

1<br />

3<br />

<br />

(D.73)<br />

(D.74)<br />

(D.75)<br />

(D.76)<br />

Unnormalized eigenvalue matrix:<br />

µ <br />

−31<br />

1 3<br />

(D.77)<br />

Normalized eigenvalue matrix:<br />

X =<br />

⎛<br />

⎜<br />

⎝<br />

−3 1<br />

√ √<br />

10 10<br />

1<br />

√<br />

10<br />

3<br />

⎞<br />

⎟<br />

⎠<br />

√<br />

10<br />

(D.78)


Appendix E 285<br />

I Final note: In practice, matrix manipulations such as matrix inversion, the evaluation<br />

of determinants, the determination of eigenvalues and eigenvectors, etc., arecarriedout<br />

numerically by using computers.<br />

I<br />

References<br />

Press 1992 W. H. Press, S. A. Teukolsky, W. T. Vetterling, B. P. Flannery, Numerical<br />

Recipes in Fortran, Cambridge University Press, Cambridge, 1992. Versions are<br />

also available for C and Pascal.


Appendix E 286<br />

Appendix E: Laplace transforms<br />

The concept of Laplace and inverse Laplace transforms is extremely useful in chemical<br />

kinetics for two reasons:<br />

(1) They connect microscopic molecular properties and statistically averaged quantities:<br />

a) () ↔ ( ): Collision cross section vs. thermal rate constant for bimolecular<br />

reactions (section ??),<br />

b) () ↔ ( ): Specific rateconstantvs. thermal rate constant for unimolecular<br />

reactions (section 8),<br />

c) () ↔ ( ): Density of states vs. partition function in statistical rate<br />

theories 8).<br />

(2) They provide a convenient method for solving of differential equations (section<br />

3.4.2).<br />

I<br />

Definition E.1: The Laplace transform L [ ()] of a function () is defined as the<br />

integral<br />

Z ∞<br />

() =L [ ()] = () − <br />

(E.1)<br />

where<br />

0<br />

• is a real variable,<br />

• () is a real function of the variable with the property () =0for 0,<br />

• is a complex variable,<br />

• () =L [ ()] is a function of the variable .<br />

I<br />

Definition E.2: The inverse Laplace transform L −1 [ ()] of the function () is<br />

defined as the integral<br />

() =L −1 [ ()] = 1 Z<br />

2<br />

+∞<br />

−∞<br />

() <br />

(E.2)<br />

where<br />

• is an arbitrary real constant.


Appendix F 287<br />

I<br />

Notes:<br />

• Since L −1 [ ()] recovers the original function (), the pair of functions ()<br />

and () is said to form a Laplace pair.<br />

• ThepropertiesoftheLaplaceandinverseLaplacetransformscanbederivedfrom<br />

those of Fourier transforms (Zachmann 1972).<br />

• Laplace transforms can be computed using symbolic algebra computer programs<br />

(Maple, Mathematica, etc.). In favorable cases, it is also possible to obtain inverse<br />

Laplace transforms by computer programs. However, it is often more convenient<br />

to take the Laplace and inverse Laplace transforms from corresponding Tables. 57<br />

I<br />

Table E.1: Table of Laplace transforms.<br />

() () =L [ ()] () () =L [ ()]<br />

()<br />

()<br />

<br />

() = R ∞<br />

0<br />

() − <br />

() − ( =0) 1<br />

− <br />

1<br />

1<br />

<br />

1<br />

( − ) 2<br />

<br />

1<br />

2<br />

sin <br />

<br />

2 + 2<br />

2 2<br />

1<br />

3<br />

cos <br />

<br />

2 + 2<br />

−1<br />

Γ ()<br />

1<br />

1 ¡<br />

− ¢ 1<br />

( − )<br />

( − )( − )<br />

Γ () is the Gamma function. For integer , Γ () =( − 1)! (see Appendix F).<br />

I<br />

References:<br />

Houston 1996 P. L. Houston, <strong>Chemical</strong> <strong>Kinetics</strong> and Reaction Dynamics, McGraw-<br />

Hill, Boston, 2001.<br />

Steinfeld 1989 J. I. Steinfeld, J. S. Francisco, W. L. Haase, <strong>Chemical</strong> <strong>Kinetics</strong> and<br />

Dynamics, Prentice Hall, Englewood Cliffs, 1989.<br />

Zachmann 1972 H. G. Zachmann, Mathematik für Chemiker, Verlag Chemie, Weinheim,<br />

1972.<br />

57 Extensive Tables of Laplace transforms are given by (Steinfeld 1989).


Appendix F 288<br />

Appendix F: The Gamma function Γ (x)<br />

• If is a positive integer, the Gamma function is<br />

Γ () =( − 1)!<br />

(F.1)<br />

• The Gamma function smoothly interpolates between all points ( ) given by<br />

=( − 1)! for positive non-integer (cf. Fig. F.1).<br />

• The Gamma function is defined for all complex numbers with a positive real<br />

part as<br />

Z ∞<br />

Γ () = −1 − <br />

(F.2)<br />

0<br />

This integral converges for complex numbers with a positive real part. It has poles<br />

for complex numbers with a negative real part (cf. Fig. F.2).<br />

I<br />

Figure F.1: The Gamma function.


Appendix G 289<br />

I<br />

Figure F.2: For negative arguments the Gamma function has poles (at all negative<br />

integers).


Appendix G 290<br />

Appendix G: Ergebnisse der statistischen Thermodynamik<br />

In diesem Kapitel sollen wichtige Ergebnisse der statistischen Thermodynamik zusammengefasst<br />

werden. Im Vordergrund steht die Berechnung der thermodynamischen Zustandsfunktionen<br />

und die Berechnung von chemischen Gleichgewichten aus molekularen<br />

Eigenschaften. Wir benutzen die Ergebnisse der Quantenmechanik:<br />

• Ein Molekül kann nur bestimmte Energiezustände einnehmen.<br />

• Diese Energiezustände sind im Prinzip nicht kontinuierlich sondern diskret verteilt.<br />

• Jeder Energiezustand hat ein bestimmtes statistisches Gewicht . Das statistische<br />

Gewicht gibt an, wie oft ein Zustand bei einer bestimmten Energie vorkommt.<br />

Beispiele:<br />

=1 eindimensionaler harmonischer Oszillator<br />

=2 isotroper zweidimensionaler harmonischer Oszillator<br />

(z.B. Biegeschwingung von CO 2 )<br />

=2 +1Rotation eines linearen Moleküls<br />

(G.1)<br />

• Die Ergebnisse der statischen Thermodynamik unterscheiden sich, je nachdem<br />

ob das betrachtete System aus unterscheidbaren oder aus ununterscheidbaren<br />

Teilchen besteht:<br />

Beispiele für unterscheidbare Teilchen: Atome im Kristallgitter<br />

Beispiele für ununterscheidbare Teilchen: 1.) Gasmoleküle<br />

2.) Elektronen (Fermi-Statistik)<br />

(G.2)<br />

• Die anzuwendende Statistik hängt vom Spin der Teilchen ab:<br />

Spin gerade: Bose-Einstein-Statistik<br />

Spin ungerade: Fermi-Dirac-Statistik<br />

(G.3)<br />

Auf die Unterschiede zwischen der Bose-Einstein- und der Fermi-Dirac-Statistik<br />

wird an anderer Stelle eingegangen (⇒ Vorlesung “Statistische Thermodynamik”).<br />

• Im Grenzfall À gilt die Boltzmann-Statistik.<br />

• Im Folgenden werden behandelt: Elektronische Anregung, Schwingungsanregung,<br />

Rotationsanregung, Translationsanregung.<br />

Die Translation erfordert abhängig vom Problem manchmal eine etwas andere<br />

Behandlung (klassische statistische Thermodynamik bzw. Semiklassik). Quantenmechanik<br />

und klassische Mechanik stimmen im Ergebnis für À überein.


Appendix G 291<br />

G.1 Boltzmann-Verteilung<br />

Die Besetzungswahrscheinlichkeit eines Zustands wird durch die Boltzmann-Verteilung<br />

angegeben. Diese ergibt sich als die wahrscheinlichste Verteilung aus dem Boltzmann-<br />

Ansatz für die Entropie<br />

= ln <br />

(G.4)<br />

wird als die thermodynamische Wahrscheinlichkeit bezeichnet.<br />

Die Boltzmann-Verteilung kann hergeleitet werden, wenn man entsprechend dem 2.<br />

Hauptsatz den Maximalwert für sucht (Methode der Lagrange’schen Multiplikatoren).<br />

Hier setzen wir die Boltzmann-Verteilung als gegeben voraus (z.B. als empirisches, experimentelles<br />

Ergebnis):<br />

I Boltzmann-Verteilung: Für diskrete Energiezustände ist die Besetzungswahrscheinlichkeit<br />

eines Zustands proportional zu − :<br />

= <br />

∝ − <br />

(G.5)<br />

Bei Berücksichtigung einer möglichen Entartung von Zuständen (Entartungsfaktor bzw.<br />

statistisches Gewicht ) und mit einer Proportionalitätskonstante 0 zur Normierung<br />

gilt<br />

= <br />

= 0 × × − <br />

(G.6)<br />

Dabei ist die Zahl der Teilchen im Energiezustand (mit dem statistischen Gewicht<br />

) und die Gesamtzahl der Teilchen.<br />

Die Funktion gibt die Wahrscheinlichkeit der Besetzung des Zustands an. Die<br />

Normierungskonstante 0 ergibt sich aus der Normierungsbedingung, dass die Gesamtwahrscheinlichkeit<br />

zu 1 herauskommt:<br />

P<br />

<br />

= X = 0 × X − =1<br />

(G.7)<br />

<br />

<br />

<br />

y 0 1<br />

= P . (G.8)<br />

− <br />

y<br />

I<br />

Normierte Boltzmann-Verteilungsfunktion:<br />

= <br />

=<br />

− <br />

P − <br />

(G.9)<br />

I Molekulare Zustandssumme: Die Summe P − im Nenner der o.a. Gleichung<br />

wird als molekulare Zustandssumme bezeichnet:<br />

= P − <br />

(G.10)<br />

Die Zustandssumme ist die zentrale Grösse der statistischen Thermodynamik. ist<br />

von der Temperatur abhängig.


Appendix G 292<br />

I<br />

Anschauliche Bedeutung der molekularen Zustandssumme:<br />

= Zahl der bei der Temperatur im Mittel besetzten Energiezustände eines Molekül.<br />

I<br />

Boltzmann-Verteilungsfunktion für kontinuierlich verteilte Energiezustände:<br />

Für kontinuierlich verteilte Energiezustände ersetzt man durch () und erhält<br />

analog eine kontinuierliche Besetzungswahrscheinlichkeit:<br />

() = ()<br />

<br />

= () × − <br />

<br />

(G.11)<br />

I Zustandsdichte: () = Zustandsdichte (=Anzahl der Zustände pro Energie-<br />

Intervall):<br />

() = ()<br />

<br />

(G.12)


Appendix G 293<br />

G.2 Mittelwerte<br />

I Mittlere Energie ²: In der Thermodynamik interessieren Mittelwerte, z.B. die mittlere<br />

Energie . Für Größen, die durch eine Verteilungsfunktion beschrieben werden,<br />

erhält man den Mittelwert nach<br />

=<br />

P<br />

P <br />

<br />

=<br />

P<br />

<br />

P<br />

<br />

=<br />

P<br />

<br />

= X P<br />

<br />

= P<br />

<br />

mit = <br />

und P =1.<br />

Übergang zu einer kontinuierlichen Verteilungsfunktion liefert<br />

Z ∞<br />

()<br />

(G.13)<br />

=<br />

0<br />

Z ∞<br />

0<br />

()<br />

(G.14)<br />

Diese Ausdrücke vereinfachen sich, wenn () auf 1 normiert ist.<br />

I<br />

Allgemeine Berechnung der Mittelwerts y einer Größe y(²):<br />

X<br />

<br />

=<br />

<br />

X<br />

<br />

<br />

(G.15)<br />

bzw.<br />

Z ∞<br />

()()<br />

=<br />

0<br />

Z ∞<br />

()<br />

(G.16)<br />

0<br />

I<br />

Example G.1: Berechnung der mittleren Energie für ein 3-Niveausystem:<br />

0 =0 1 = 1 2 =2 1<br />

0 =3 2 =2 2 =1<br />

= X =6<br />

(G.17)<br />

(G.18)


Appendix G 294<br />

Ergebnis:<br />

=<br />

1 X<br />

= X <br />

<br />

= X <br />

(G.19)<br />

<br />

<br />

<br />

= 1 6 (3 0 +2 1 +1 2 ) (G.20)<br />

= 1 6 (3 0 +2 1 +2 1 ) (G.21)<br />

= 1 6 (0 + 4 1) (G.22)<br />

= 4 6 1 (G.23)<br />

= 2 3 1<br />

(G.24)<br />

¤<br />

I Mittlere Energie eines Moleküls: Bei Gültigkeit der Boltzmann-Verteilung kann die<br />

mittlere Energie eines Moleküle aus der molekularen Zustandssumme berechnet werden:<br />

=<br />

=<br />

1 X<br />

= X <br />

<br />

<br />

P<br />

− <br />

P<br />

− <br />

<br />

= X <br />

<br />

(G.25)<br />

(G.26)<br />

y<br />

=<br />

P<br />

− <br />

= 2<br />

<br />

= 2<br />

<br />

= 2<br />

<br />

= 2<br />

<br />

<br />

X<br />

<br />

X<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

2 − <br />

|× 2<br />

2<br />

(G.27)<br />

(G.28)<br />

¡<br />

<br />

− <br />

<br />

|(durch differenzieren zeigen) (G.29)<br />

X<br />

− <br />

(G.30)<br />

<br />

= 2 ln <br />

<br />

(G.31)<br />

(G.32)


Appendix G 295<br />

G.3 Anwendung auf ein Zweiniveau-System<br />

I<br />

Chemisch relevante Beispiele für Zweiniveau-Systeme:<br />

• Paramagnetische Atome u. Moleküle<br />

• NMR-Spektroskopie: Atomkerne mit Kernspin. Typischer Frequenzbereich:<br />

100 MHz.<br />

• Molekül mit cis-trans-Isomeren oder 2 Konformeren<br />

I<br />

Figure G.1: Zwei-Niveau-System: Energiezustände<br />

The Two Level System: Energy Levels<br />

<br />

<br />

<br />

<br />

g <br />

<br />

g <br />

<br />

Zustandssumme:<br />

Q <br />

i<br />

0 / kT 1<br />

/ kT<br />

1e<br />

1e<br />

1e<br />

i<br />

/ kT<br />

i<br />

g e<br />

1<br />

/ kT<br />

Besetzungsverhältnis:<br />

0/<br />

kT<br />

0 g0e<br />

N<br />

<br />

N<br />

N<br />

N<br />

Q<br />

1<br />

/ kT<br />

1 g1e<br />

N<br />

<br />

N<br />

<br />

Q<br />

1<br />

/ kT<br />

1 g1e<br />

<br />

0<br />

/ kT<br />

0 ge 0<br />

e<br />

1<br />

/ kT<br />

<strong>Physical</strong> <strong>Chemistry</strong> III: <strong>Chemical</strong> <strong>Kinetics</strong> • © F. Temps, IPC Kiel 173<br />

I<br />

Besetzungsverhältnis:<br />

mit<br />

1<br />

= 1 − 1<br />

<br />

(G.33)<br />

= 0 − 0 + 1 − 1<br />

=1+ − 1<br />

(G.34)<br />

(G.35)<br />

(1) Grenzfall für → 0:<br />

1<br />

→∞y − 1 → −∞ =0<br />

y Bei → 0 ist nur der tiefste Energiezustand besetzt.<br />

(G.36)


Appendix G 296<br />

(2) Grenzfall für →∞:<br />

1<br />

→ 0 y − 1 → −0 =1<br />

(G.37)<br />

y Bei →∞ergibt sich eine Gleichbesetzung der Energiezustände. Dies ist die<br />

maximale Besetzung im oberen Zustand!<br />

• Eine Besetzungsinversion ( 1 0 ) würde formal eine negative Temperatur<br />

erfordern (→ Laser).<br />

I<br />

Figure G.2: Zwei-Niveau-System: Boltzmann-Besetzung der Zustände (rot: unterer<br />

Zustand, blau: oberer Zustand).<br />

The Two Level System: Boltzmann Distribution<br />

N i<br />

N<br />

1<br />

= 1000 J/mol<br />

0.75<br />

0.5<br />

0.25<br />

0<br />

0<br />

250<br />

500<br />

750<br />

1000<br />

T /(K)<br />

<strong>Physical</strong> <strong>Chemistry</strong> III: <strong>Chemical</strong> <strong>Kinetics</strong> • © F. Temps, IPC Kiel 174<br />

I<br />

Mittlere thermische Energie:<br />

(1) 1 =125kJ mol (entsprechend =1000cm −1 ; z.B. C-C-Schwingung im IR<br />

oder Energieabstand zwischen cis-trans-Isomeren):<br />

½<br />

1<br />

= − 1 8 × 10 =<br />

−3 @ = 300K<br />

(G.38)<br />

0 023 @ =1000K<br />

(2) 1 =004 kJ mol (entsprechend Energieabstand zwischen Kernspinzuständen bei<br />

100 MHz NMR):<br />

1<br />

0<br />

= − 1 =0999984 @ =300K<br />

(G.39)<br />

Reihenentwicklung:<br />

−1 =1− 1<br />

+ ≈ 1 − 1<br />

(G.40)<br />

<br />

y sehr kleiner Besetzungsunterschied y Wunsch der NMR-Spektroskopiker zu<br />

immer größeren Frequenzen (Rekord heute 950 MHz).


Appendix G 297<br />

I<br />

Figure G.3: Zwei-Niveau-System: Mittlere thermische Energie.<br />

The Two Level System: Energy<br />

E<br />

1000<br />

= 1000 J/mol<br />

750<br />

500<br />

250<br />

0<br />

0<br />

250<br />

500<br />

750<br />

1000<br />

T /(K)<br />

<strong>Physical</strong> <strong>Chemistry</strong> III: <strong>Chemical</strong> <strong>Kinetics</strong> • © F. Temps, IPC Kiel 175<br />

I<br />

Spezifische Wärme:<br />

= <br />

<br />

(G.41)<br />

I<br />

Figure G.4: Zwei-Niveau-System: Spezifische Wärme.<br />

The Two Level System: Specific Heat<br />

c v<br />

5<br />

= 1000 J/mol<br />

4<br />

3<br />

2<br />

1<br />

0<br />

0<br />

250<br />

500<br />

750<br />

1000<br />

T /(K)<br />

<strong>Physical</strong> <strong>Chemistry</strong> III: <strong>Chemical</strong> <strong>Kinetics</strong> • © F. Temps, IPC Kiel 176


Appendix G 298<br />

G.4 Mikrokanonische und makrokanonische Ensembles<br />

Die statistische Thermodynamik macht Aussagen über statistische Ensembles<br />

(Ansammlungen von Systemen, Molekülen). Wir betrachten verschiedene Ensembles<br />

von miteinander wechselwirkenden Molekülen.<br />

I Der Begriff des Ensembles: Wir betrachten ein abgeschlossenes System bei konstantem<br />

Volumen, konstanter Zusammensetzung und konstanter Temperatur (bzw. bestimmter<br />

Energie (s.u.)). Ein Ensemble erhalten wir, wenn wir uns N Replikationen<br />

dieser Systeme denken, die wir unter Einhaltung bestimmter Regeln erhalten (Regel =<br />

Canon). Alle Systeme in dem Ensemble sollen im thermischen Gleichgewicht miteinander<br />

(sie dürfen zwischen einander Energie austauschen).<br />

Es werden verschiedene Sorten von Ensembles unterschieden:<br />

• Mikrokanonisches Ensemble: Ensemble von Systemen (z.B. Atomen,<br />

Molekülen) mit konstanter Teilchenzahl, konstantem Volumen, konstanter Energie.<br />

• Makrokanonisches Ensemble (kurz kanonisches Ensemble genannt): Ensemble<br />

von Systemen (z.B. Atomen, Molekülen) mit konstanter Teilchenzahl, konstantem<br />

Volumen, konstanter Temperatur.<br />

• Großkanonisches Ensemble: Ensemble von offenen Systemen (z.B. Teilchen,<br />

Molekülen) mit konstantem Volumen, konstanter Temperatur, die zwischeneinander<br />

Teilchen austauschen können.<br />

I Systemzustandssumme (= kanonische Zustandssumme): Molekulare Zustandssumme:<br />

= P alle Molekülzustände − <br />

(G.42)<br />

Systemzustandssumme (= kanonische Zustandssumme):<br />

= P alle Systemzustände − <br />

(G.43)<br />

Innere Energie des Systems:<br />

= 2 ln <br />

<br />

(G.44)<br />

Bei mehreren Teilchensorten (A, B, C, . . . ): Zustandsummen sind multiplikativ,<br />

da die Energien additiv sind, y<br />

= × × × <br />

(G.45)


Appendix G 299<br />

• Systemzustandssumme für unterscheidbare Teilchen der Sorte (z.B.<br />

Kristall):<br />

Y <br />

= = <br />

<br />

(G.46)<br />

y für 1mol:<br />

= 2 ln <br />

<br />

<br />

2<br />

ln <br />

= <br />

<br />

(G.47)<br />

• Systemzustandssumme für nicht unterscheidbare Teilchen der Sorte (z.B.<br />

Gas):<br />

y für 1mol ( = ):<br />

= <br />

<br />

!<br />

(G.48)<br />

= 2 ln <br />

<br />

= 2 <br />

ln <br />

!<br />

(G.49)


Appendix G 300<br />

G.5 Statistische Interpretation der thermodynamischen Zustandsgrößen<br />

I<br />

Innere Energie:<br />

= 2 ln <br />

<br />

(G.50)<br />

I<br />

Helmholtz-Energie:<br />

<br />

= − <br />

(G.51)<br />

y<br />

y<br />

<br />

<br />

µ <br />

<br />

<br />

<br />

=<br />

<br />

µ <br />

<br />

<br />

2<br />

= − <br />

2<br />

<br />

<br />

− <br />

−<br />

µ <br />

<br />

<br />

− <br />

= − ln <br />

= −<br />

<br />

2<br />

<br />

= − ln <br />

(G.54)<br />

= − ln <br />

(G.52)<br />

(G.53)<br />

(G.55)<br />

I<br />

Entropie:<br />

= −<br />

µ <br />

<br />

<br />

= − <br />

<br />

= (G.56)<br />

I<br />

Druck:<br />

= −<br />

µ <br />

<br />

<br />

µ ln <br />

= <br />

<br />

<br />

= (G.57)


Appendix G 301<br />

G.6 Chemisches Gleichgewicht<br />

y<br />

bzw.<br />

+ À <br />

=exp<br />

µ− ∆ <br />

ª<br />

<br />

∆ ª = − ln <br />

(G.58)<br />

(G.59)<br />

(G.60)<br />

I Gleichgewichtskonstante: Wir berechnen zunächst<br />

= − ln ,<br />

(G.61)<br />

= <br />

<br />

indem wir <br />

!<br />

ln − anwenden:<br />

etc.. . . einsetzen und die Stirling’sche Näherung ln ! =<br />

= − [ln +ln +ln ]<br />

"<br />

#<br />

= − ln <br />

<br />

<br />

<br />

! +ln <br />

! +ln <br />

!<br />

= − [( ln − ln !)<br />

+( ln − ln !)<br />

+( ln − ln !)]<br />

= − [( ln − ln + )<br />

+( ln − ln + )<br />

+( ln − ln + )]<br />

(G.62)<br />

(G.63)<br />

(G.64)<br />

(G.65)<br />

I Gleichgewichtsbedingung: Im chemischen Gleichgewicht muss gelten<br />

µ <br />

=0 (G.66)<br />

<br />

y<br />

0= [( ln − ln + )<br />

<br />

+( ln − ln + )<br />

+( ln − ln + )]<br />

(G.67)<br />

µ<br />

<br />

1<br />

= ln − − ln +1<br />

<br />

µ<br />

<br />

<br />

1<br />

+ ln − − ln +1<br />

<br />

µ<br />

<br />

<br />

1<br />

− ln − − ln +1<br />

(G.68)<br />

<br />

=(ln − ln )+(ln − ln ) − (ln − ln ) (G.69)<br />

=ln <br />

<br />

− ln <br />

<br />

(G.70)


Appendix G 302<br />

y<br />

<br />

<br />

=<br />

<br />

<br />

(G.71)<br />

Übergang auf Teilchendichten und volumenbezogenen Zustandssummen:<br />

y<br />

( )<br />

( )( ) = ( )<br />

( )( )<br />

=<br />

( )<br />

( )( ) = ∗ <br />

∗ ∗ <br />

(G.72)<br />

(G.73)<br />

Bezug auf die getrennten molekularen Nullpunktsenergien:<br />

=<br />

0 <br />

0 0 <br />

× −∆ 0 <br />

(G.74)<br />

mit 0 = die auf das Volumen bezogene Molekülzustandssumme:<br />

0 = <br />

<br />

× × × (G.75)<br />

I Endergebnis: Die Indices ∗ und 0 wurden hier zur Klarheit benutzt. Der Einfachheit<br />

halber werden diese üblicherweise weggelassen.<br />

y<br />

=<br />

<br />

<br />

× −∆ 0 <br />

(G.76)


Appendix G 303<br />

G.7 Zustandssumme für elektronische Zustände<br />

Die elektronische Zustandssumme muss unter Einschluss aller elektronischen Entartungen<br />

explizit berechnet werden:<br />

= X <br />

− <br />

(G.77)<br />

Speziell sind zu berücksichtigen:<br />

• Spinentartung (Spinmultiplizität ):<br />

=2 +1<br />

(G.78)<br />

• elektronischer Bahndrehimpuls ( (Atome) bzw. Λ (lin. Moleküle) bzw. Symmetrierasse<br />

oder für nichtlin. Moleküle):<br />

• elektronische Feinstrukturkomponenten (Index im Atom-Termsymbol):<br />

=2 +1<br />

(G.79)<br />

Nur niedrig liegende Elektronenzustände liefern einen Beitrag. Die Zustandsumme ist<br />

explizit auszurechnen.


Appendix G 304<br />

G.8 Zustandssumme für die Schwingungsbewegung<br />

I<br />

Harmonischer Oszillator:<br />

= =0 1 2 (G.80)<br />

I<br />

Schwingungszustandssumme:<br />

∞X<br />

= − =<br />

y<br />

=0<br />

∞X<br />

=0<br />

−·<br />

mit = <br />

<br />

(G.81)<br />

= © 1+ − + −2 + −3 + ª mit = − 1 (G.82)<br />

=1+ + 2 + 3 + <br />

(G.83)<br />

= 1 für 1<br />

1 − <br />

(G.84)<br />

=<br />

∙ µ<br />

1<br />

1 − = 1 − exp − <br />

¸−1<br />

− <br />

(G.85)<br />

I<br />

s Oszillatoren:<br />

=<br />

Y<br />

=1<br />

∙ µ<br />

1 − exp − ¸−1<br />

<br />

<br />

(G.86)<br />

I<br />

typische Werte bei Zimmertemperatur:<br />

≈ 1 10<br />

(G.87)<br />

I Grenzwert µ von Q für T →∞bzw. für ν → 0: Für →∞bzw. für → 0<br />

kann exp − <br />

entwickelt werden:<br />

<br />

µ<br />

exp − <br />

≈ 1 − <br />

+ <br />

(G.88)<br />

y<br />

1<br />

1<br />

= µ<br />

1 − exp − = µ<br />

1 − 1 − (G.89)<br />

<br />

+ <br />

y<br />

≈ <br />

(G.90)<br />

<br />

Für klassische Oszillatoren:<br />

≈<br />

Y<br />

=1<br />

<br />

<br />

(G.91)


Appendix G 305<br />

I<br />

Abweichungen vom harmonischen Oszillator:<br />

• Für anharmonische Oszillatoren ist die Zustandssumme am besten durch explizite<br />

numerische Berechnung nach P ∞<br />

=0 − zu ermitteln.<br />

• Torsionsschwingungen können bei tiefen Temperaturen als harmonische Oszillatoren<br />

beschrieben werden, bei hohen Temperaturen als freie 1D-Rotatoren. Im<br />

Zwischenbereich muss die Zustandssumme ebenfalls explizit berechnet werden<br />

(wobei die Energiezustände für den gehinderten 1D-Rotator zugrundegelegt werden<br />

müssen).


Appendix G 306<br />

G.9 Zustandssumme für die Rotationsbewegung<br />

I<br />

lineare Moleküle:<br />

= ~2<br />

( +1) =0 1 2 (G.92)<br />

2<br />

~ 2<br />

2 = <br />

(G.93)<br />

I<br />

Rotationszustandssumme:<br />

(2) = 1 <br />

= 1 <br />

= 1 <br />

∞X<br />

− <br />

(G.94)<br />

=0<br />

∞X<br />

(2 +1) −~2 (+1)2<br />

=0<br />

Z ∞<br />

(G.95)<br />

(2 +1) −~2 (+1)2 (G.96)<br />

=0<br />

y<br />

(2) = 2<br />

~ 2 = <br />

<br />

(G.97)<br />

I<br />

Symmetriezahl:<br />

<br />

(G.98)<br />

I Rotationszustandssumme für nichtlineare Moleküle: ist nur für symmetrische<br />

Kreisel analytisch darstellbar, nicht für asymmetrische Kreisel. Für nicht zu tiefe Temperaturen<br />

ist<br />

(3) = 12<br />

<br />

µ 2<br />

~ 2 32<br />

( ) 12 = 12<br />

()32 () −12 (G.99)<br />

I<br />

typische Werte bei Zimmertemperatur:<br />

≈ 100 1000<br />

(G.100)


Appendix G 307<br />

G.10 Zustandssumme für die Translationsbewegung<br />

I<br />

Teilchen im 1D-Kasten:<br />

³<br />

= 2 <br />

´2<br />

8 × =1 2 3 (G.101)<br />

<br />

I<br />

1D-Translationszustandssumme:<br />

y<br />

=<br />

=<br />

∞X<br />

− =<br />

<br />

Z ∞<br />

1<br />

−2 2 8 2 <br />

(G.102)<br />

µ 2<br />

2 12<br />

× ¡ in cm −1¢ (G.103)<br />

<br />

<br />

= µ 2<br />

2 12<br />

(G.104)<br />

I<br />

Teilchen im 3D-Kasten:<br />

" µ 2 µ 2 µ # 2<br />

= 2<br />

8 × <br />

+ +<br />

<br />

(G.105)<br />

I<br />

3DTranslationszustandssumme:<br />

y<br />

µ 32 2<br />

=<br />

× <br />

<br />

<br />

2<br />

=<br />

¡<br />

in cm<br />

−3 ¢ (G.106)<br />

µ 2<br />

2 32<br />

(G.107)<br />

I<br />

typische Werte:<br />

≈ 10 24 × <br />

(G.108)<br />

y<br />

<br />

<br />

≈ 10 24 cm −3<br />

(G.109)

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!