You are on page 1of 68

Theories of Molecular Reaction

Dynamics: The Microscopic Foundation


of Chemical Kinetics 2nd Edition Niels
E. Henriksen
Visit to download the full and correct content document:
https://ebookmass.com/product/theories-of-molecular-reaction-dynamics-the-microsc
opic-foundation-of-chemical-kinetics-2nd-edition-niels-e-henriksen/
T H E O R I E S O F M O L E C U L A R R E AC T I O N DY N A M I C S
Theories of Molecular Reaction Dynamics
The Microscopic Foundation of Chemical Kinetics

Second Edition

Niels Engholm Henriksen and


Flemming Yssing Hansen
Department of Chemistry, Technical University of Denmark

1
3
Great Clarendon Street, Oxford, OX2 6DP,
United Kingdom
Oxford University Press is a department of the University of Oxford.
It furthers the University’s objective of excellence in research, scholarship,
and education by publishing worldwide. Oxford is a registered trade mark of
Oxford University Press in the UK and in certain other countries
© Niels Engholm Henriksen and Flemming Yssing Hansen 2019
The moral rights of the authors have been asserted
First Edition published in 2008
First Edition published in paperback in 2012
Second Edition published in 2019
Impression: 1
All rights reserved. No part of this publication may be reproduced, stored in
a retrieval system, or transmitted, in any form or by any means, without the
prior permission in writing of Oxford University Press, or as expressly permitted
by law, by licence or under terms agreed with the appropriate reprographics
rights organization. Enquiries concerning reproduction outside the scope of the
above should be sent to the Rights Department, Oxford University Press, at the
address above
You must not circulate this work in any other form
and you must impose this same condition on any acquirer
Published in the United States of America by Oxford University Press
198 Madison Avenue, New York, NY 10016, United States of America
British Library Cataloguing in Publication Data
Data available
Library of Congress Control Number: 2018947008
ISBN 978–0–19–880501–4
DOI: 10.1093/oso/9780198805014.001.0001
Printed and bound by
CPI Group (UK) Ltd, Croydon, CR0 4YY
Links to third party websites are provided by Oxford in good faith and
for information only. Oxford disclaims any responsibility for the materials
contained in any third party website referenced in this work.
Preface to the First Edition

This book focuses on the basic concepts in molecular reaction dynamics, which is the
microscopic atomic-level description of chemical reactions, in contrast to the macro-
scopic phenomenological description known from chemical kinetics. It is a very extensive
field and we have obviously not been able, or even tried, to make a comprehensive
treatment of all contributions to this field. Instead, we limited ourselves to give a
reasonable coherent and systematic presentation of what we find to be central and
important theoretical concepts and developments, which should be useful for students at
the graduate or senior undergraduate level and for researchers who want to enter the field.
The purpose of the book is to bring about a deeper understanding of the atomic
processes involved in chemical reactions and to show how rate constants may be
determined from first principles. For example, we show how the thermally averaged rate
constant k(T ), known from chemical kinetics, for a bimolecular gas-phase reaction may
be calculated as proper averages of rate constants for processes that are highly specified
in terms of the quantum states of reactants and products, and how these state-to-state
rate constants can be related to the underlying molecular dynamics. The entire spectrum
of elementary reactions, from isolated gas-phase reactions, such as in molecular-beam
experiments, to condensed-phase reactions, are considered. Although the emphasis
has been on the development of analytical theories and results that describe essential
features in a chemical reaction, we have also included some aspects of computational
and numerical techniques that are used when the simpler analytical results are no longer
accurate enough.
We have tried, without being overly formalistic, to develop the subject in a systematic
manner with attention to basic concepts and clarity of derivations. The reader is assumed
to be familiar with the basic concepts of classical mechanics, quantum mechanics,
and chemical kinetics. In addition, some knowledge of statistical mechanics is required
and, since not all potential readers may have that, we have included an Appendix
that summarizes the most important results of relevance. The book is reasonably
self-contained such that a standard background in mathematics, physics, and physical
chemistry should be sufficient and make it possible for the students to follow and
understand the derivations and developments in the book. A few sections may be a little
more demanding, in particular some of the sections on quantum dynamics and stochastic
dynamics.
Earlier versions of the book have been used in our course on advanced physical
chemistry and we thank the students for many useful comments. We also thank our
colleagues, in particular Dr. Klaus B. Møller for making valuable contributions and
comments.
vi Preface to the First Edition

The book is divided into three parts. Chapters 2–8 are on gas-phase reactions,
Chapters 9–11 on condensed-phase reactions, and Appendices A–I contain details about
concepts and derivations that were not included in the main body of the text. We have
put a frame around equations that express central results to make it easier for the reader
to navigate among the many equations in the text.
In Chapter 2, we develop the connection between the microscopic description of
isolated bimolecular collisions and the macroscopic rate constant. That is, the reaction
cross-sections that can be measured in molecular-beam experiments are defined and the
relation to k(T ) is established. Chapters 3 and 4 continue with the theoretical microscopic
description of isolated bimolecular collisions. Chapter 3 has a description of potential
energy surfaces, that is, the energy landscapes for the nuclear dynamics. Potential energy
surfaces are first discussed on a qualitative level. The more quantitative description of
the energetics of bond breaking and bond making is considered, where this is possible
without extensive numerical calculations, leading to a semi-analytical result in the form
of the London equation. These considerations cannot, of course, replace the extensive
numerical calculations that are required in order to obtain high quality potential energy
surfaces. Chapter 4 is the longest chapter of the book with the focus on the key issue
of the nuclear dynamics of bimolecular reactions. The dynamics is described by the quasi-
classical approach as well as by exact quantum mechanics, with emphasis on the relation
between the dynamics and the reaction cross-sections.
In Chapter 5, attention is directed toward the direct calculation of k(T ), that is, a
method that bypasses the detailed state-to-state reaction cross-sections. In this approach
the rate constant is calculated from the reactive flux of population across a dividing
surface on the potential energy surface, an approach that also prepares for subsequent
applications to condensed-phase reaction dynamics. In Chapter 6, we continue with the
direct calculation of k(T ) and the whole chapter is devoted to the approximate but very
important approach of transition-state theory. The underlying assumptions of this theory
imply that rate constants can be obtained from a stationary equilibrium flux without any
explicit consideration of the reaction dynamics.
In Chapter 7, we turn to the other basic type of elementary reaction, that is,
unimolecular reactions, and discuss detailed reaction dynamics as well as transition-state
theory for unimolecular reactions. In this chapter, we also touch upon the question of the
atomic-level detection and control of molecular dynamics. In the final chapter dealing with
gas-phase reactions, Chapter 8, we consider unimolecular as well as bimolecular reactions
and summarize the insights obtained concerning the microscopic interpretation of the
Arrhenius parameters, that is, the pre-exponential factor and the activation energy of the
Arrhenius equation.
Chapters 9–11 deal with elementary reactions in condensed phases. Chapter 9 is on
the energetics of solvation and, for bimolecular reactions, the important interplay between
diffusion and chemical reaction. Chapter 10 is on the calculation of reaction rates according
to transition-state theory, including static solvent effects that are taken into account via the
so-called potential-of-mean force. Finally, in Chapter 11, we describe how dynamical
effects of the solvent may influence the rate constant, starting with Kramers theory and
continuing with the more recent Grote–Hynes theory for k(T ). Both theories are based
Preface to the First Edition vii

on a stochastic dynamical description of the influence of the solvent molecules on the


reaction dynamics.
We have added several appendices that give a short introduction to important disci-
plines such as statistical mechanics and stochastic dynamics, as well as developing more
technical aspects like various coordinate transformations. Furthermore, examples and
end-of-chapter problems illustrate the theory and its connection to chemical problems.
Preface to the Second Edition

The first edition of this book was published about ten years ago. The present edition
reflects our efforts to refine and improve clarity of material already in the first edition,
and to elaborate the treatment of various topics. The aim of the book is unchanged with
an emphasis on basic concepts and the development of insights provided by analytical
results.
We have, in particular, included new material/sections concerning: (i) adiabatic
and non-adiabatic electron-nuclear dynamics, (ii) classical two-body models of chem-
ical reactions, for example, the so-called Langevin model for ion-molecule reactions,
(iii) quantum mechanical models for crossing of one-dimensional barriers, and (iv) a
more detailed description of the Born and Onsager models for solvation.
A major rewrite of some sections or whole chapters include Sections 5.1 and 7.5 as well
as Chapters 10 and 11. In addition, there are several small changes spread throughout
the book and new end-of-chapter problems have been added, as well as a new appendix.
Finally, a remark concerning notation. For notational convenience (following many
other textbooks) we have often omitted the integration limits of definite (multidimen-
sional-)integrals when it is understood that integration is over “all space.”
1
Introduction

Chemical reactions, the transformation of matter at the atomic level, are distinctive
features of chemistry. They include a series of basic processes from the transfer of
single electrons or protons to the transfer of groups of nuclei and electrons between
molecules, that is, the breaking and formation of chemical bonds. These processes are of
fundamental importance to all aspects of life in the sense that they determine the function
and evolution in chemical and biological systems.
The transformation from reactants to products can be described at either a phe-
nomenological level, as in classical chemical kinetics, or at a detailed molecular level,
as in molecular reaction dynamics.1 The former description is based on experimental
observation and, combined with chemical intuition, rate laws are proposed to enable a
calculation of the rate of the reaction. It does not provide direct insight into the process at
a microscopic molecular level. The aim of molecular reaction dynamics is to provide such
an insight as well as to deduce rate laws and calculate rate constants from basic molecular
properties and dynamics. Dynamics is, in this context, the description of atomic motion
under the influence of a force or, equivalently, a potential.
The main objectives of molecular reaction dynamics may be briefly summarized by
the following points:

• the microscopic foundation of chemical kinetics;


• state-to-state chemistry and chemistry in real time;
• control of chemical reactions at the microscopic level.

Before we go on and discuss these objectives in more detail, it might be appropriate to


consider the relation between molecular reaction dynamics and the science of physical
chemistry. Normally, physical chemistry is divided into four major branches, as sketched
in the figure that follows (each of these areas are based on fundamental axioms). At
the macroscopic level, we have the old disciplines: “thermodynamics” and “kinetics.”
At the microscopic level we have “quantum mechanics,” and the connection between
the two levels is provided by “statistical mechanics.” Molecular reaction dynamics

1 The roots of molecular reaction dynamics go back to a famous paper by H. Eyring and M. Polanyi, Z. Phys.
Chem. B12, 279 (1931).

Theories of Molecular Reaction Dynamics. Second Edition. Niels E. Henriksen and Flemming Y. Hansen, Oxford University
Press 2019. © Niels E. Henriksen and Flemming Y. Hansen. DOI: 10.1093/oso/9780198805014.001.0001
2 Introduction

encompasses (as sketched by the oval) the central branches of physical chemistry, with
the exception of thermodynamics.

Thermodynamics Kinetics

cs
mi
yna
Statistical mechanics

on di
eact
la r r
lecu
Quantum mechanics

Mo
Interaction between
matter and light

A few concepts from classical chemical kinetics should be recalled [1]. Chemical
change is represented by a reaction scheme. For example,

2H2 + O2 → 2H2 O

The rate of reaction, R, is the rate of change in the concentration of one of the reactants
or products, such as R = −d[H2 ]/dt, and the rate law giving the relation between the rate
and the concentrations can be established experimentally.
This reaction scheme represents, apparently, a simple reaction but it does not proceed
as written. That is, the oxidation of hydrogen does not happen in a collision between two
H2 molecules and one O2 molecule. This is also clear when it is remembered that all the
stoichiometric coefficients in such a scheme can be multiplied by an arbitrary constant
without changing the content of the reaction scheme. Thus, most reaction schemes show
merely the overall transformation from reactants to products without specifying the path
taken. The actual path of the reaction involves the formation of intermediate species
and includes several elementary steps. These steps are known as elementary reactions and
together they constitute what is called the reaction mechanism of the reaction. It is a great
challenge in chemical kinetics to discover the reaction mechanism, that is, to unravel
which elementary reactions are involved.
Elementary reactions are reactions that directly express basic chemical events, that is,
the making or breaking of chemical bonds. In the gas phase, there are only two types of
elementary reactions:2

2 The existence of trimolecular reactions is sometimes suggested. For example, H + OH + M → H O + M,


2
where M is a third body. However, the reaction probably proceeds by a two-step mechanism,
Introduction 3

• unimolecular reaction (e.g., due to the absorption of electromagnetic radiation);


• bimolecular reaction (due to a collision between two molecules);

and in condensed phases, in addition, a third type:

• bimolecular association/recombination reaction.

The reaction schemes of elementary reactions are to be taken literally. For example, one
of the elementary reactions in the reaction between hydrogen and oxygen is a simple
atom transfer:

H + O2 → OH + O

In this bimolecular reaction the stoichiometric coefficients are equal to one, meaning that
one hydrogen atom collides with one oxygen molecule. Once the reaction mechanism
and all the rate constants for the elementary reactions are known, the reaction rates for
all species are given by a simple set of coupled first-order differential equations. These
equations can be solved quite easily on a computer, and give the concentrations of all
species as a function of time. These results may then be compared with experimen-
tal results.
From this discussion, it follows that: elementary reactions are at the heart of chemistry.
The study of these reactions is the main subject of this book.
The rearrangement of nuclei in an elementary chemical reaction takes place over a
distance of a few ångström (1 ångström = 10−10 m) and within a time of about 10–100
femtoseconds (1 femtosecond = 10−15 s; a femtosecond is to a second what one second
is to 32 million years!), equivalent to atomic speeds of the order of 1 km/s. The challenges
in molecular reaction dynamics are: (i) to understand and follow in real time the detailed
atomic dynamics involved in the elementary processes, (ii) to use this knowledge in the
control of these reactions at the microscopic level, for example, by means of external
laser fields, and (iii) to establish the relation between such microscopic processes and
macroscopic quantities like the rate constants of the elementary processes.
We consider the detailed evolution of isolated elementary reactions3 in the gas phase,
for example,

A + BC(n) −→ AB(m) + C

At the fundamental level, the course of such a reaction between an atom A and a
diatomic molecule BC is governed by quantum mechanics. Thus, within this theoretical

i.e., (1) H + OH → H2 O∗ , and (2) H2 O∗ + M → H2 O + M, where H2 O∗ is an energy-rich water molecule


with an energy that exceeds the dissociation limit, and the function of M is to take away the energy. That is, the
reaction actually proceeds via bimolecular collisions.
3 An elementary reaction is defined as a reaction that takes place as written in the reaction scheme. We will
here distinguish between a truly elementary reaction, where the reaction takes place in isolation without any
secondary collisions, and the traditional definition of an elementary reaction, where inelastic collisions among
the molecules in the reaction scheme (or with container walls) can take place.
4 Introduction

framework the reaction dynamics at a given collision energy can be analyzed for reactants
in a given quantum state (denoted by the quantum number n) and one can extract the
transition probability for the formation of products in various quantum states (denoted
by the quantum number m). At this level, we consider the “state-to-state” dynamics of
the reaction.
When we consider elementary reactions, it should be realized that the outcome of a
bimolecular collision can also be non-reactive. Thus,

A + BC(n) −→ BC(m) + A

and we distinguish between an elastic collision process, if quantum states n and m


are identical, and otherwise, an inelastic collision process. Note that inelastic collisions
correspond to energy transfer between molecules—in the present case, for example, the
transfer of relative translational energy between A and BC to vibrational energy in BC.
The realization of an isolated elementary reaction is experimentally difficult. The
closest realization is achieved under the highly specialized laboratory conditions of an
ultra-high vacuum molecular-beam experiment. Most often collisions between molecules
in the gas phase occur, making it impossible to obtain state-to-state specific information
because of the energy exchange in such collisions. Instead, thermally averaged rate
constants may be obtained. Thus, energy transfer, that is, inelastic collisions among
the reactants, implies that an equilibrium Boltzmann distribution is established for the
collision energies and over the internal quantum states of the reactants. A parameter in
the equilibrium Boltzmann distribution is the macroscopic temperature T . Under such
conditions, the well-known rate constant k(T ) of chemical kinetics can be defined and
evaluated based on the underlying detailed dynamics of the reaction.
The macroscopic rate of reaction is, typically, much slower than the rate that can
be inferred from the time it takes to cross the transition states (i.e., all the intermediate
configurations between reactants and products) because the fraction of reactants with
sufficient energy to react is very small at typical temperatures.
Reactions in a condensed phase are never isolated but under strong influence of the
surrounding solvent molecules. The solvent will modify the interaction between the
reactants, and it can act as an energy source or sink. Under such conditions the state-
to-state dynamics described here cannot be studied, and the focus is then turned to the
evaluation of the rate constant k(T ) for elementary reactions. The elementary reactions in
a solvent include both unimolecular and bimolecular reactions as in the gas phase and, in
addition, bimolecular association/recombination reactions. That is, an elementary reaction
of the type A + BC → ABC, which can take place because the products may not fly
apart as they do in the gas phase. This happens when the products are not able to escape
from the solvent “cage” and the ABC molecule is stabilized due to energy transfer to the
solvent.4 Note that one sometimes distinguishes between association as an outcome of a
bimolecular reaction and recombination as the inverse of unimolecular fragmentation.

4 Association/recombination can, under special conditions, also take place in the gas phase (in a single
elementary reaction step), e.g., in the form of so-called radiative recombination; see Section 6.5.
Nuclear Dynamics: The Schrödinger Equation 5

On the experimental side, the chemical dynamics on the state-to-state level is being
studied via molecular-beam and laser techniques [2]. Alternative and complementary
techniques have been developed in order to study the real-time evolution of elementary
reactions [3]. Thus, the time resolution in the observation of chemical reactions has
increased dramatically over the last decades. The “race against time” has recently
reached the ultimate femtosecond resolution with the direct observation of chemical
reactions as they proceed along the reaction path via transition states from reactants
to products. This spectacular achievement was made possible by the development of
femtosecond lasers, that is, laser pulses with a duration as short as a few femtoseconds.
In a typical experiment two laser pulses are used, a “pump pulse” and a “probe pulse.”
The first femtosecond pulse initiates a chemical reaction, say the breaking of a chemical
bond in a unimolecular reaction, and a second time-delayed femtosecond pulse probe
this process. The ultrashort duration of the pump pulse implies that the zero of time is
well defined. The probe pulse is, for example, tuned to be in resonance with a particular
transition in one of the fragments and, when it is fired at a series of time delays relative
to the pump pulse, one can directly observe the formation of the fragment. This type of
real-time chemistry is called femtosecond chemistry (or simply, femtochemistry). Another
interesting aspect of femtosecond chemistry concerns the challenging objective of using
femtosecond lasers to control the outcome of chemical reactions, say to break a particular
bond in a large molecule. This type of control at the molecular level is much more
selective than traditional methods for control where only macroscopic parameters like the
temperature can be varied. In short, femtochemistry is about the detection and control of
transition states, that is, the intermediate short-lived states on the path from reactants to
products.
On the theoretical side, advances have also been made both in methodology and
in concepts. For example, new and powerful techniques for the solution of the time-
dependent Schrödinger equation (see Section 1.1) have been developed. New concepts
for laser control of chemical reactions have been introduced where, for example, one
laser pulse can create a non-stationary nuclear state that can be intercepted or redirected
with a second laser pulse at a precisely timed delay.
The theoretical foundation for reaction dynamics is quantum mechanics and statistical
mechanics. In addition, in the description of nuclear motion, concepts from classical
mechanics play an important role. A few results of molecular quantum mechanics and
statistical mechanics are summarized in Sections 1.1 and 1.2. In the second part of the
book, we will return to concepts and results of particular relevance to condensed-phase
dynamics.

1.1 Nuclear Dynamics: The Schrödinger Equation


The reader is assumed to be familiar with some of the basic concepts of quantum
mechanics. At this point we will therefore just briefly consider a few central concepts,
including the time-dependent Schrödinger equation for nuclear dynamics. This equation
allows us to focus on the nuclear motion associated with a chemical reaction.
6 Introduction

We consider a system of K electrons and N nuclei, interacting through Coulomb


forces. The basic equation of motion in quantum mechanics, the time-dependent
Schrödinger equation, can be written in the form

∂(rlab , Rlab , t)
i h̄ = (T̂ nuc + Ĥ e )(rlab , Rlab , t) (1.1)
∂t

where i is the imaginary unit, h̄ = h/(2π ) is the Planck constant divided by 2π , and
the wave function depends on rlab = (r1 , r2 , . . . , rK ) and Rlab = (R1 , R2 , . . . , RN ), which
denote all electron and nuclear coordinates, respectively, measured relative to a fixed
laboratory coordinate system. The operators are


N
P̂ 2g
T̂ nuc = (1.2)
2Mg
g=1

which is the kinetic energy operator of the nuclei, where P̂ g = −i h̄∇g is the momentum
operator and Mg the mass of the gth nucleus, and Ĥ e is the so-called electronic
Hamiltonian including the internuclear repulsion,


K
p̂2i 
N 
K
Zg e2 
K 
K
e2 
N 
N
Zg Zh e2
Ĥ e = − + + (1.3)
2me 4π 0 rig 4π 0 rij 4π 0 rgh
i=1 g=1 i=1 i=1 j>i g=1 h>g

where the first term represents the kinetic energy of the electrons, the second term the
attraction between electrons and nuclei, the third term the electron–electron repulsion,
and the last term the internuclear repulsion. More specifically, p̂i = −i h̄∇i is the momen-
tum operator of the ith electron and me its mass, 0 in the Coulomb terms is the vacuum
permittivity, rig is the distance between electron i and nucleus g, the other distances rij and
rgh have a similar meaning, and Zg e is the electric charge of the gth nucleus, where Zg
is the atomic number. The Hamiltonian is written in its non-relativistic form, that is,
spin-orbit terms, and so on are neglected. Note that the electronic Hamiltonian does not
depend on the absolute positions of the nuclei but only on internuclear distances and the
distances between electrons and nuclei.
The translational motion of the particles as a whole (i.e., the center-of-mass motion)
can be separated out. This is done by a change of variables from rlab , Rlab to RCM and
r, R, where RCM gives the position of the center of mass and r, R are internal coordinates
that describe the relative position of the electrons with respect to the nuclei and the
relative position of the nuclei, respectively. This coordinate transformation implies

(rlab , Rlab , t) = (RCM , t)(r, R, t) (1.4)

where (RCM , t) is the wave function associated with the free translational motion of the
center of mass, and (r, R, t) describes the internal motion, given by a time-dependent
Nuclear Dynamics: The Schrödinger Equation 7

Schrödinger equation similar to Eq. (1.1). The kinetic energy operators expressed in the
internal coordinates take, however, a more complicated form than specified here, (see,
e.g., Appendix E).5
Fortunately, a direct solution of Eq. (1.1) is normally not necessary. The electrons
are very light particles whereas the nuclei are, at least, about three orders of magnitude
heavier. From the point of view of the electronic state, the nuclear positions can be
considered as slowly changing external parameters, which means that the electrons
experience a slowly changing potential. To that end, it is convenient to expand the total
wave function for the nuclear and electronic degrees of freedom in the form

(r, R, t) = χi (R, t)ψi (r; R) (1.5)
i

where ψi (r; R) is the usual stationary electronic wave function determined as an eigen-
function of the electronic Hamiltonian

Ĥ e ψi (r; R) = Ei (R)ψi (r; R) (1.6)

In this electronic Schrödinger equation, the nuclei are fixed and the equation has
only a parametric dependence on the nuclear coordinates, as indicated by the “;” in
the electronic wave function, Ei (R) is the corresponding electronic energy (including
internuclear repulsion), which is a function of the nuclear geometry. Equation (1.6) is
solved for different fixed values of the nuclear coordinates such that the wave function
and energy are evaluated over the entire range of relevant R values. χi (R, t) in Eq. (1.5) is
the wave function for the nuclear motion associated with the ith electronic state, and for
fixed values of the nuclear coordinates, it can be considered as an “expansion coefficient”
in an expansion that, in principle, is exact for a complete set of electronic states.
The expansion in Eq. (1.5) is motivated by the anticipation that the electrons expe-
rience a slowly changing potential due to the nuclear motion. To that end, when the
electrons are in a given quantum state (say the electronic ground state) it can be shown
that the quantum number of the electronic state, in the following indicated by the
subscript i, is unchanged as long as the nuclear motion can be considered as being
sufficiently slow. Thus, no transitions among the electronic states will take place under
these conditions. This is the physical basis for the so-called adiabatic approximation, which
can be written in the form,

adia (r, R, t) = χ(R, t)ψi (r; R) (1.7)

corresponding to a single term in Eq. (1.5), and for notational convenience we have
dropped the subscript on the nuclear wave function.

5 Normally, three approximations are introduced in this context: (i) the center of mass is taken to be identical
to the center of mass of the nuclei; (ii) the kinetic energy operators of the electrons are taken to be identical
to the expression given in Eq. (1.3), which again means that the nuclei are considered to be infinitely heavy
compared to the electrons; and (iii) coupling terms between the kinetic energy operators of the electrons and
nuclei, introduced by the transformation, are neglected.
8 Introduction

The solution to Eq. (1.6) for the electronic energy in, for example, a diatomic molecule
is well known. In this case there is only one internuclear coordinate and the electronic
energy, Ei (R), is consequently represented by a curve as a function of the internuclear
distance. For small displacements around the equilibrium bond length R = R0 , the curve
can be represented by a quadratic function. Thus, when we expand to second order
around a minimum at R = R0 ,
   2 
∂E ∂ E
E(R) = E(R0 ) + (R − R0 ) + (1/2) (R − R0 )2 + · · ·
∂R 0 ∂R2 0
= E(R0 ) + (1/2)k(R − R0 )2 + · · · (1.8)

where the subscript indicates that the derivatives are evaluated at the minimum, and
 
∂ 2E
k= (1.9)
∂R2 0

is the force constant. However, when we consider chemical reactions, where chemical
bonds are formed and broken, the electronic energy for all internuclear distances is
important. The description of the simultaneous making and breaking of chemical bonds
leads to multidimensional potential energy surfaces that are discussed in Chapter 3.
Substituting Eq. (1.7) into Eq. (1.1), we obtain (see Appendix A)

∂χ(R, t)
i h̄ = (T̂ nuc + Ei (R) + ψi |T̂ nuc |ψi 0 )χ(R, t) (1.10)
∂t

where we have used that ψi |∇g |ψi  = ∇g ψi |ψi /2 = 0, when ψi (r; R) is real, and the
electronic wave function is normalized. The subscript on the matrix element implies
that T̂ nuc acts only on ψi and the matrix element involves an integration over electron
coordinates.
The term ψi |T̂ nuc |ψi 0 is normally very small compared to the electronic energy,
and may consequently be dropped (the resulting approximation is often referred to as
the Born–Oppenheimer approximation):

∂χ(R, t)
i h̄ = [T̂ nuc + Ei (R)]χ(R, t) (1.11)
∂t

Equation (1.11) is the fundamental equation of motion within the adiabatic approxima-
tion. We see that: the nuclei move on a potential energy surface given by the electronic energy.
Thus, the electron-nuclear dynamics has been separated and one must first solve for
the electronic energy, Eq. (1.6), and subsequently solve the time-dependent Schrödinger
equation for the nuclear motion, Eq. (1.11).
The physical implication of the adiabatic approximation is that the electrons remain
in a given electronic eigenstate during the nuclear motion. The electrons follow the
Nuclear Dynamics: The Schrödinger Equation 9

A+B
C
RAB
3
Before 2.5
0.5 2
1 1.5
1.5
2 1
RBC 2.5 C
3 0.5 +
A+B
C AB
RAB
3
After 2.5
0.5 2
1 1.5
1.5
2
RBC 2.5
1
3 0.5 +C
AB

Fig. 1.1.1 Schematic illustration of the probability density, |χ (R, t)|2 , associated with a chemical
reaction, A + BC → AB + C. The contour lines represent the potential energy surface (see Chapter 3),
and the probability density is shown at two times: before the reaction where only reactants are present,
and after the reaction where products as well as reactants are present. The arrows indicate the direction
of motion associated with the relative motion of reactants and products. (Note that, due to the finite
uncertainty in the A–B distance, RAB , there is some uncertainty in the initial relative translational energy
of A + BC.)

nuclei, for example, as a reaction proceeds from reactants to products, such that the
electronic state “deforms” in a continuous way without electronic transitions. From a
more practical point of view, the approximation implies that we can separate the solutions
to the electronic and nuclear motion.
The probability density, |χ(R, t)|2 , may be used to calculate the reaction probability.
The probability density associated with the nuclear motion of a chemical reaction  is
illustrated in Fig. 1.1.1. The reaction probability may be evaluated from P = R∈Prod
|χ(R, t ∼ ∞)|2 dR, where the integration is restricted to configurations representing
the products (Prod), in the example for large B–C distances, RBC . The limit t ∼ ∞
implies that the probability density obtained long after reaction is used in the integral.
In this limit, where the reaction is completed, there is a negligible probability density
in the region where RAB as well as RBC are small. In practice, “long after” is identified as
the time where the reaction probability becomes independent of time and, typically, this
situation is established after a few hundred femtoseconds.
The general time-dependent solutions to Eq. (1.11) are denoted as non-stationary
states. They can be expanded in terms of the eigenstates φn (R) of the Hamiltonian

Ĥ = T̂ nuc + Ei (R) (1.12)

with φn (R) given by

Ĥ φn (R) = En φn (R) (1.13)


10 Introduction

where n is a set of quantum numbers that fully specify the eigenstates of the nuclear
Hamiltonian. The general solution can then be written in the form6 (as can be checked
by direct substitution into Eq. (1.11))

χ(R, t) = exp(−i Ĥ t/h̄)χ(R, 0)



= exp(−i Ĥ t/h̄) cn φn (R)
n

−i En t/h̄
= cn φn (R)e (1.14)
n

Each state in the sum, n (R, t) = φn (R)e−i En t/h̄ , is denoted as a stationary state, because
ˆ n (t) (e.g., for the operator representing the position)
all expectation values  n (t)|A|
are independent of time. That is, there is no observable time dependence associated with
a single stationary state. Equation (1.14) shows that the non-stationary time-dependent
solutions can be written as a superposition of the stationary solutions, with coefficients
that are independent of time. The coefficients are determined by the way the system was
prepared at t = 0.
The eigenstates of Ĥ are well known, for non-interacting molecules, say the reactants
A + BC (an atom and a diatomic molecule), giving quantized vibrational and rotational
energy levels. Within the so-called rigid-rotor approximation where couplings between
rotation and vibration are neglected, Eq. (1.12) can for non-interacting molecules be
written in the form

Ĥ 0 = Ĥ trans + Ĥ vib + Ĥ rot (1.15)

where Ĥ trans represents the free relative motion of A and BC, and Ĥ vib and Ĥ rot corre-
spond to the vibration and rotation of BC, respectively. This form of the Hamiltonian,
with a sum of independent terms, implies that the eigenstates take the form

J
φn0 (R) = φtrans (Rrel )φvib
n
(R)φrot (θ , φ) (1.16)

where the functions in the product are eigenfunctions corresponding to translation,


vibration, and rotation. The eigenvalues are

En0 = Etr + En + EJ (1.17)

That is, the total energy is the sum of the energies associated with translation, vibration,
and rotation. The translational energy is continuous (as in classical mechanics).

6 A function of an operator is defined through its (Taylor) power series. The summation sign should really
be understood as a summation over discrete quantum numbers and an integration over continuous labels
corresponding to translational motion.
Nuclear Dynamics: The Schrödinger Equation 11

E4

E3

E2

E1

E0

Fig. 1.1.2 The energy levels of a one-dimensional


√ harmonic oscillator.The zero-point energy E0 = h̄ω/2,
where for a diatomic molecule ω = k/μ, k is the force constant, and μ is the reduced mass.

For a one-dimensional harmonic oscillator, with the potential in Eq. (1.8), the Ham-
iltonian is

h̄2 ∂ 2
Ĥ vib = − + (1/2)k(R − R0 )2 (1.18)
2μ ∂R2

where μ is the reduced mass of the two nuclei in BC. This Hamiltonian has the well-
known eigenvalues

En = h̄ω(n + 1/2) , n = 0, 1, 2, . . . (1.19)


where ω = 2π ν = k/μ, and ν is identical to the frequency of the corresponding classical
harmonic motion. The vibrational energy levels of the one-dimensional harmonic
oscillator are illustrated in Fig. 1.1.2. Note, for example, that the vibrational zero-point
energy changes under isotope substitution since the reduced mass will change. We shall
see, later on, that this purely quantum mechanical effect can change the magnitude of
the macroscopic rate constant k(T ).
The rigid-rotor Hamiltonian for a diatomic molecule with the moment of inertia
I = μR02 is

L̂ 2
Ĥ rot = (1.20)
2I
12 Introduction

where L̂ is the angular momentum operator, giving the well-known eigenvalues

EJ = h̄2 J(J + 1)/(2I ) , J = 0, 1, 2, . . . (1.21)

with the degeneracy ωJ = 2J + 1.


Typically, the energy spacing between rotational energy levels is much smaller than
the energy spacing between vibrational energy levels that, in turn, is much smaller
than the energy spacing between electronic energy levels:

Erot Evib Eelec (1.22)

where the energy spacing is defined as the energy difference between adjacent energy
levels.
It is possible to solve Eq. (1.11) numerically for the nuclear motion associated with
chemical reactions and to calculate the reaction probability including detailed state-
to-state reaction probabilities (see Section 4.2). However, with the present computer
technology such an approach is in practice limited to systems with a small number of
degrees of freedom.
For practical reasons, a quasi-classical approximation to the quantum dynamics
described by Eq. (1.11) is often sought. In the quasi-classical trajectory approach
(discussed in Section 4.1) only one aspect of the quantum nature of the process is
incorporated in the calculation: the initial conditions for the trajectories are sampled in
accord with the quantized vibrational and rotational energy levels of the reactants.
Obviously, purely quantum mechanical effects cannot be described when one replaces
the time evolution by classical mechanics. Thus, the quasi-classical trajectory approach
exhibits, for example, the following deficiencies: (i) zero-point energies are not con-
served properly (they can, e.g., be converted to translational energy) and (ii) quantum
mechanical tunneling cannot be described.
Finally, it should be noted that the motion of the nuclei is not always confined to a
single electronic state (as assumed in Eq. (1.7)). This situation can, for example, occur
when two potential energy surfaces come close together for some nuclear geometry. The
dynamics of such processes are referred to as non-adiabatic. When several electronic
states are in play, Eq. (1.11) must be replaced by a matrix equation with a dimension
given by the number of electronic states (see Section 4.2). The equation contains
coupling terms between the electronic states, implying that the nuclear motion in all
the electronic states is coupled.

1.2 Thermal Equilibrium: The Boltzmann Distribution


Statistical mechanics gives the relation between microscopic information such as quan-
tum mechanical energy levels and macroscopic properties. Some important statistical
mechanical concepts and results are summarized in Appendix B. Here we will briefly
review one central result: the Boltzmann distribution for thermal equilibrium.
Thermal Equilibrium: The Boltzmann Distribution 13

For reactants in complete thermal equilibrium, the probability of finding a BC


molecule in a specific quantum state, n, is given by the Boltzmann distribution (see
Appendix B.1). Thus, in the special case of non-interacting molecules the probability,
pBC(n) , of finding a BC molecule in the internal (electronic, vibrational, and rotational)
quantum states with energy En is

ωn
pBC(n) = exp(−En /kB T ) (1.23)
QBC

where ωn is the degeneracy of the nth quantum level (i.e., the number of states with the
same energy En ) and QBC is the “internal” partition function of the BC molecule where
center-of-mass motion is excluded, given by

QBC = ωn exp(−En /kB T ) (1.24)
n

that is, a weighted sum over all energy levels, where the weights are proportional to the
occupation probabilities of each level.
The distribution depends on the temperature; only the lowest energy level is populated
at T = 0. When the temperature is raised, higher energy levels will also be populated. The
probability of populating high energy levels decreases exponentially with the energy.
The Boltzmann distribution is illustrated in Fig. 1.2.1 for the vibrational states of
a one-dimensional harmonic oscillator with the frequency ω = 2π ν, where the energy
levels are given by Eq. (1.19), and in Fig. 1.2.2 for the rotational states of a linear molecule

E4

E3

E2

E1

E0
pn

Fig. 1.2.1 The Boltzmann distribution for a system with equally spaced energy levels En and identical
degeneracy ωn of all levels (T > 0). This figure gives the population of states at the temperature T for a
harmonic oscillator.
14 Introduction

E5

E0 pJ

√ distribution for the rotational energy of a linear molecule (T > 0). The
Fig. 1.2.2 The Boltzmann
maximum is at Jmax = IkB T/h̄ − 1/2 (rounded off to the closest integer).

with the moment of inertia I , where the energy levels are given by Eq. (1.21) with the
degeneracy ωJ = 2J + 1.

P(E)

kBT/2 E

Fig. 1.2.3 The Boltzmann distribution for free translational motion. The maximum is at Emax =
kB T/2.

The Boltzmann distribution for free translational motion takes a special form (see
Appendix B.2.1), since the energy is continuous in this case. The probability of finding
a translational energy in the range Etr , Etr + dEtr is given by
 3/2 
1
P(Etr )dEtr = 2π Etr exp(−Etr /kB T )dEtr (1.25)
π kB T

This function is shown in Fig. 1.2.3.


Problems 15

These distribution functions show that, at a given temperature, many energy levels
will be populated when the energy spacing is small (which is the case for translational
and rotational degrees of freedom), whereas only a few of the lowest energy levels will
have a substantial population when the energy spacing is large (vibrational and electronic
degrees of freedom). Furthermore, it is often found that the degeneracy of the energy
levels is increasing as a function of the energy. The maxima of the distributions will
therefore often be found for energies above the ground-state energy. Thus, for the
translational energy, the maximum is at the energy kB T /2.
An elementary reaction is, in classical chemical kinetics, defined under conditions where
energy transfer among the molecules in the reaction scheme or with surrounding solvent
molecules can take place. In this case, we write

A + BC −→ AB + C

for an elementary reaction. We have deleted the quantum numbers associated with the
molecules, and it is understood that the states are populated according to the Boltzmann
distribution. Furthermore, when the reaction takes place, we will normally assume that
the thermal equilibrium among the reactants can be maintained at all times.

Further reading/references
[1] J.I. Steinfeld, J.S. Francisco, and W.L. Hase, Chemical kinetics and dynamics, second edition
(Prentice Hall, 1999).
[2] D.R. Herschbach, Angewandte Chemie—international edition in English 26, 1221 (1987).
Y.T. Lee, Science 236, 793 (1987). J.C. Polanyi, Science 236, 680 (1987).
[3] A.H. Zewail, Scientific American, Dec. 1990, page 40. A.H. Zewail, J. Phys. Chem. 104, 5660
(2000).
[4] R.D. Levine, Molecular reaction dynamics (Cambridge University Press, 2005).

.....................................................................................................................................

P RO B L E M S

1.1 Consider a Hamiltonian that can be written in the form

Ĥ = Ĥ 1 + Ĥ 2 + · · ·

where Ĥ 1 , Ĥ 2 , · · · refers to different independent sets of coordinates. Assume that


the eigenfunctions and associated eigenvalues are known for each Hamiltonian, that
is Ĥ i ψi = Ei ψi , for i = 1, 2, · · · . Show that

ψ = ψ1 ψ2 · · ·
16 Introduction

is an eigenfunction of Ĥ with associated eigenvalue

E = E1 + E2 + · · ·

This important result was used in connection with Eq. (1.15) and is used several
times throughout the book.
1.2 Show that Eq. (1.14) is a solution to Eq. (1.11).
1.3 Show that the maximum in the √ Boltzmann distribution for the rotational energy of a
linear molecule is at Jmax = IkB T /h̄ − 1/2 (when J is considered as a continuous
variable).
1.4 Consider the Boltzmann distribution for free translational motion, and calculate the
probability of finding translational energies that exceed E = E ∗ . Compare with the
expression exp[−E ∗ /(kB T )].
∞
Use the integral: √4π x u2 e−u du = 1 − erf(x) + √2xπ e−x ,
2 2

x
where the error function is defined as: erf(x) = √2π 0 e−u du.
2

Note that erf(0) = 0, erf(0.5) = 0.5205, erf(1.0) = 0.8427, erf(2.0) = 0.9953,


and erf(∞) = 1.
1.5 Elementary concepts of probability and statistics play an important role in this book.
Thus, these concepts are an integral part of, for example, quantum mechanics
and statistical mechanics. The probability that some continuous variable x lies
between x and x + dx is denoted by P(x)dx. Often we refer to P(x) as the probability
distribution for x (although P(x) strictly speaking is a probability density).
The average value or mean value of a variable x, which can take any value between
−∞ and ∞, is defined by
 ∞
x = xP(x)dx
−∞

(a) Calculate the average translational energy using Eq. (1.25).


The variance of x is defined by

σx2 = (x − x)2 

where σx is called the standard deviation. It is a measure of the spread of the


distribution about its mean value.
(b) Show that σx2 = x2  − x2 , where x2  is the average value of x2 .
(c) Calculate the standard deviation associated with the Boltzmann distribution for
translational motion, Eq. (1.25).
Problems 17

In connection with the evaluation of the integrals the Gamma function (n) is useful.
It is defined as
 ∞
(n) = xn−1 exp(−x)dx, n > 0
0

with the special values

(n + 1) = n!, where n = 1, 2, . . .,


1 · 3 · 5 · · · (2n − 1) √
(n + 1/2) = π , where n = 1, 2, . . ..
2n
2
From Microscopic to
Macroscopic Descriptions

Key ideas and results


In this chapter, we consider bimolecular reactions from both a microscopic and
a macroscopic point of view and thereby derive a theoretical expression for the
macroscopic phenomenological rate constant. That is, a relation between molecular
reaction dynamics and chemical kinetics is established.
The outcome of an isolated (microscopic) reactive scattering event can be specified
in terms of an intrinsic fundamental quantity: the reaction cross-section. The cross-
section is an effective area that the reactants present to each other in the scattering
process. It depends on the quantum states of the molecules as well as the relative
speed of the reactants, and it can be calculated from the collision dynamics (to be
described in Chapter 4).
In this chapter, we define the cross-section and derive its relation to the rate
constant. We show the following.
• The macroscopic rate constant is related to the relative speed of the reactants and
the reaction cross-section, and the expression contains a weighted average over
all possible quantum states and velocities of the reactants, and sums and integrals
over all possible quantum states and velocities of the products.
• Specialized to thermal equilibrium, the velocity distributions for the molecules
are the Maxwell–Boltzmann distribution (a special case of the general Boltzmann
distribution law). The expression for the rate constant at temperature T , k(T ),
can be reduced to an integral over the relative speed of the reactants. Also, as
a consequence of the time-reversal symmetry of the Schrödinger equation, the
ratio of the rate constants for the forward and the reverse reaction is equal to the
equilibrium constant (detailed balance).

In chemical kinetics, we learn that an elementary bimolecular reaction,

A + B −→ products (2.1)

Theories of Molecular Reaction Dynamics. Second Edition. Niels E. Henriksen and Flemming Y. Hansen, Oxford University
Press 2019. © Niels E. Henriksen and Flemming Y. Hansen. DOI: 10.1093/oso/9780198805014.001.0001
22 From Microscopic to Macroscopic Descriptions

obeys a second-order rate law, given by

d[A]
− = k[A][B] (2.2)
dt

where k ≡ k(T ) is the temperature-dependent bimolecular rate constant. The purpose


of the following chapters (Chapters 2–6) is to obtain an in-depth understanding of this
relation and the factors that determine k(T ).

2.1 Cross-Sections and Rate Constants


We begin by establishing the relation between the so-called reaction cross-section σR and
the bimolecular rate constant. Let us consider an elementary gas-phase reaction,

A(i, vA ) + B(j, vB ) → C(l, vC ) + D(m, vD ) (2.3)

where an A and a B molecule collide,1 and a C and D molecule are formed. The reactant
molecule A is prior to the collision in a given internal quantum state i, which specifies a
set of quantum numbers corresponding to the rotational, vibrational, and electronic state
of the molecule, and moves with velocity vA relative to some laboratory fixed coordinate
system (the velocity is specified by a vector with a given direction and length, |vA |, which
is the speed). Reactant molecule B is likewise in a given internal quantum state j and
moves with velocity vB . The product molecules move with velocities vC and vD , and end
up in internal quantum states as specified by the quantum numbers l and m, respectively.
These conditions are readily specified in a theoretical calculation of the reaction but
difficult to realize in an experiment, because inelastic molecular collisions will upset the
detailed specification of the molecular states. The requirements of an experimental set-
up for the investigation of the chemical reaction in Eq. (2.3) may be summarized in the
following way.

• Establishment and maintenance of two molecular beams, where the molecules move
in a specified direction with a specified speed and are in a specified internal quantum
state.
• Detection of internal quantum states, direction of motion, and speed of product
molecules after the collision.
• Single-collision conditions, that is, there is one and just one collision in the reaction
zone defined as the zone where the beams cross, and no collisions prior to or after
this collision.

1 The word “collision” should not be taken too literally, since molecules are not, say, hard spheres where
it is straightforward to count the “hits.” Thus, a “collision” should really be interpreted as the broader term
“a scattering event.”
Cross-Sections and Rate Constants 23

A(i) dΩ
V
vA

C(l,vc)

B(j)

vB

Fig. 2.1.1 Idealized molecular-beam experiment for the reaction A(i, vA ) + B(j, vB ) → C(l, vC ) +
D(m, vD ). The coordinate system is fixed in the laboratory. The reactants move with the relative speed
v = |vA − vB |.

These requirements can be met in a so-called crossed molecular-beam experiment, which


is sketched in Fig. 2.1.1. Here we can generate beams of molecules with well-defined
velocities and it is possible to determine the speed of the product molecules, for example,
vC = |vC |, by the so-called time-of-flight technique. The elimination of multiple scattering
in the reaction zone and collisions in the beams are obtained by doing the experiments
in high vacuum, that is, at very low pressures.
In an experiment, we can monitor the number of product molecules, C or D, emerging
in a space angle d around the direction ; d is given by the physical design of the
detector and  by its position ( is conveniently specified by the two polar angles θ
and φ). This is the simplest analysis of a scattering process, where we just count the
number of product molecules independent of their internal state and speed. In a more
advanced analysis, one may use the time-of-flight technique to analyze the speed of
product molecules, and only monitor products with a certain speed, that is, C molecules
with a speed in the range vC , vC + dvC and D molecules with a speed in the range
vD , vD + dvD . A still more advanced detection method also allows for the detection of
the quantum states of the product molecules. This is the ultimate degree of specification
of a scattering experiment.
At sufficiently low pressure (as in the beam experiment) where an A molecule
only collides with one single B molecule in the reaction zone, it will hold that the
number of product molecules is proportional to the number of collisions between A and
B molecules. Clearly, that number depends on the relative speed of the two molecules,
v = |vA − vB |, the time interval dt, and the number of B molecules. Therefore, if we
assume that the number density (number/m3 ) of B molecules in quantum state j and with
velocity vB is nB(j,vB ) , and that the flux density of A molecules relative to the B molecules
24 From Microscopic to Macroscopic Descriptions

is JA(i,v) (number/(m2 s)), then the number of collisions between A and B in the time
interval dt is proportional to nB(j,vB ) VJA(i,v) Adt. Here, V is the volume of the reaction
zone (see Fig. 2.1.1), and A the cross-sectional area of the beam of A molecules.
In the experiment sketched in Fig. 2.1.1, we monitor the number of product molecules,
C(l, vC ), emerging in the space angle d around the direction , with the speed in
the range vC , vC + dvC , and in the internal quantum state specified by l. Further-
more, let us for the moment assume that we can also detect the state of the product
molecule D, m, vD , as specified in Eq. (2.3). The number of product molecules,
dNC(l,vC ,vC +dvC ) (,  + d, t, t + dt)m,v , registered in the detector in the time interval
D
dt, given that D is in the state m, vD , may therefore be written as (using that the change
in the value of a function, can be written on the form df = (df /dx)dx)

 3N (, 
d ) t) 
dNC(l,vC ,vC +dvC ) (,  + d, t, t + dt)m,v ≡
C(l,v C
 dvC ddt
D dvC ddt  (2.4)
m,vD
= P (ij, v|ml, vC , , vD ; t)nB(j,vB ) JA(i,v) V A dvC ddt

where we have introduced the factor P (ij, v|ml, vC , , vD ; t), which is proportional to
the probability that a product molecule is formed and found in the specified state. Note
that on the left-hand side of the vertical bar in the argument list we have written the
quantum numbers and the relative speed that specifies the state of the reactants, whereas
the quantum numbers and velocities on the right-hand side specify the state of the
products. The notation |m,vD implies that the number of C molecules in the specified
state is counted only when D is in the state m, vD .
The complete degree of specification of a scattering experiment is rarely realized in
an actual experiment and, normally, we will just monitor the number of C molecules in
the specified state, irrespective of the quantum state and velocity of D. In order to obtain
that quantity, we integrate over vD and sum over m in Eq. (2.4). Thus,

d 3 NC(l,vC ) (, t)
dvC ddt = P (ij, v|l, vC , ; t)nB(j,vB ) JA(i,v) V A dvC ddt (2.5)
dvC ddt
 
where P (ij, v|l, vC , ; t) = m all vD P (ij, v|ml, vC , , vD ; t)dvD . By division with VdvC
ddt, we obtain an expression for the number of product molecules per reaction zone
volume, per space angle, per time unit, and per unit speed, d 3 nC(l,vC ) /(dvC ddt), where
nC(l,vC ) = NC(l,vC ) /V is the number density of product molecules in the reaction zone.
We find

d 3 nC(l,vC ) (, t)
= P (ij, v|l, vC , ; t)A nB( j,vB ) JA(i,v) (2.6)
dvC ddt

The time dependence of the number of particles in the third-order differential and in
the probability can be dropped, since the experiments are typically conducted under
stationary conditions. Thus,
Cross-Sections and Rate Constants 25

d 3 nC(l,vC ) ()
= P (ij, v|l, vC , )A nB(j,vB ) JA(i,v)
dvC ddt
 2  (2.7)
d σR
≡ (ij, v|l, vC , )nB(j,vB ) JA(i,v)
dvC d

 2 
where we have replaced PA by dvd Cσd R
(ij, v|l, vC , ), which defines the differential
reactive scattering cross-section. It has the dimension of an area per unit speed, per unit
space angle, because P is a probability density, that is, P dvC d is dimensionless. Since
nB(j,vB ) , JA(i,v) , and d 3 nC(l) /(dvC ddt) are intensive properties (i.e., independent of the
size of the system), PA and d 2 σR /(dvC d) must therefore also be intensive properties
independent of the beam geometries. The differential cross-section is a function of the
quantum states (ijl), the relative speed v = |vA − vB | of the reactants, and the continuous
velocity of the product, specified by the space angle  and the speed vC . Since it is
an intensive property of the chemical reaction, it is often used to report the results of
scattering experiments. Physically, the cross-section represents an effective area that the
reactants present to each other in connection with a scattering process.
We now need expressions for the relative flux of A molecules with regard to the B
molecules, and the number of B molecules. Let us introduce the following notation that
allows for any distribution of velocities in the molecular beams:

number density of A(i) with velocity in the range vA , vA + dvA : nA(i) fA(i) (vA )dvA
number density of B(j) with velocity in the range vB , vB + dvB : nB(j) fB(j) (vB )dvB

where nA(i) is the number density (number/m3 ) of A in quantum state i, irrespective


of their velocity, and fA(i) (vA )dvA is the normalized velocity probability distribution of
A(i), that is, the probability of finding an A(i) with velocity in the range from vA to
vA + dvA , and nB(j) is the number density (number/m3 ) of B molecules in quantum state
j. The product of the incoming flux, JA(i,v) , and the number of B molecules may then be
expressed as

JA(i,v) nB(j,vB ) = vnA(i) fA(i) (vA )dvA × nB(j) fB(j) (vB )dvB (2.8)

and we obtain from Eq. (2.7) the following expression:

 
d 3 nC(l,vC ) () d 2 σR
= (ij, v|l, vC , ) vnA(i) fA(i) (vA )nB(j) fB(j) (vB )dvA dvB (2.9)
dvC ddt dvC d

Note that the idealized beam experiment corresponds to “sharp” distributions in


the velocities, which mathematically can be expressed in terms of the so-called delta
26 From Microscopic to Macroscopic Descriptions

function2 δ(x). For example, if the velocity of all the A molecules is v0A , we have
fA(i) (vA ) = δ(vA − v0A ).
Often we are not interested in the distribution of speeds of the product molecules, so
we may accordingly multiply both sides of the equation by dvC and integrate over the
speed:
 
d 2 nC(l) () dσR
= (ij, v|l, ) vnA(i) fA(i) (vA )nB(j) fB(j) (vB )dvA dvB (2.10)
ddt d
    2 
∞ d σR
where dσ d = 0
R
dvC d dvC .
Likewise, if we multiply both sides of the equation by d and integrate over all , we
get an expression for the total reaction rate of the state-to-state reaction as specified by
the internal quantum numbers and the relative speed of the reactants:

dnC(l)
= σR (ij, v|l)vfA(i) (vA )fB(j) (vB )dvA dvB nA(i) nB(j) (2.11)
dt

where
4π  
dσR
σR (ij, v|l) = (ij, v|l, )d (2.12)
0 d

is the integrated cross-section, often referred to as the total cross-section for the
state-to-state reaction as specified by the internal quantum numbers.3 Since the
cross-section is still resolved with respect to the quantum states, it is also referred to as
a partial cross-section. The various cross-sections are summarized in the following table:

Reaction cross-section Dimension


 2 
d σR
Differential (ij, v|l, vC , ) area/(speed × space angle)
dvC d
 
dσR
Differential (ij, v|l, ) area/(space angle)
d

Partial σR (ij, v|l) area

Equation (2.11) has the form of a normal rate equation for the isolated bimolecular
reaction given by (compare with Eq. (2.3))

2 The delta function has the property


∞  
−∞ δ(x − x )f (x)dx = f (x ), where f (x) is an arbitrary function.
Additional properties of the delta function are described, e.g., in many textbooks on quantum mechanics.
3 The 4π indicates that integration is over the full unit sphere, d = sin θ dθ dφ, with θ ∈ [0, π ] and φ ∈
[0, 2π ].
Cross-Sections and Rate Constants 27

A(i, vA ) + B(j, vB ) → C(l) + D (2.13)

This rate, at the state-to-state level, can be measured provided that all inelastic collisions
are eliminated. Note that the rate constant (with the proper unit for a bimolecular
reaction, volume/time) can be identified as the cross-section times the relative speed
σR (ij, v|l)v.
At this point, let us consider how to interpret the cross-section. Since the partial cross-
section is an intensive property, it is clear from Eqs (2.7) and (2.10) that P is an extensive
variable given by

dσR (ij, v|l, )/d


P (ij, v|l, ) = (2.14)
A

or, after integration over d, Ptot = P (ij, v|l) = σR (ij, v|l)/A.
That is, the reaction probability is proportional to the ratio of the reaction cross-
section and the area of the beam (A); see Fig. 2.1.2.

Example 2.1 Molecular-beam studies, some experimental data

Many elementary chemical reactions have been investigated via molecular-beam techniques.
An example is the reaction
F + H2 → HF + H
and its variant with D2 [see D.M. Neumark, A.M. Wodtke, G.N. Robinson, C.C. Hayden,
and Y.T. Lee, J. Chem. Phys. 82, 3045 (1985) and M. Faubel, L. Rusin, S. Schlemmer,
F. Sondermann, U. Tappe, and J.P. Toennies, J. Chem. Phys. 101, 2106 (1994)]. It is found
that the total reaction cross-section increases with collision energy. Differential cross-sections
associated with angular distributions of products were resolved with respect to the different
vibrational states of HF(n). The angular distributions of HF(n) depend on the vibrational state
and were all found to be non-isotropic.
The formation of the stronger HF bond, with a bond dissociation energy that is about
130 kJ/mol higher than for H2 , implies a substantial drop in potential energy and hence a
large release of kinetic energy that can be distributed among the translational, vibrational, and
rotational degrees of freedom of the products. At collision energies from 2.9 to 14.2 kJ/mol, it
is found that the HF vibrational distribution is highly inverted, with most of the population in
n = 2 and n = 3.
Another reaction studied via molecular-beam techniques is the (SN 2) reaction:
Cl− + CH3 Br → ClCH3 + Br−
where the total cross-section as a function of the relative collision energy has been determined
[L.A. Angel and K.M. Ervin, J. Am. Chem. Soc. 125, 1014 (2003)]. A special feature of this
(ion–molecule) reaction is found at low collision energies. Thus, with increasing collision
energies over the range 0.06–0.6 eV, the cross-section declines from 1.3 × 10−16 cm2 to
0.08 × 10−16 cm2 .
28 From Microscopic to Macroscopic Descriptions

Fig. 2.1.2 A beam of molecules incident on a rectangle of area A. The ratio of the total cross-section
(σR (ij, v|l)) to the area of the rectangle (A) is, according to Eq. (2.14), related to the fraction of molecules
that are undergoing reaction.

The total rate of reaction, dnC /dt = −dnA /dt, is obtained from Eq. (2.11) by summing
over all possible quantum states of reactants and products and all possible velocities vA
and vB . We find

− dnA /dt = vσR (ij, v|l)fA(i) (vA )fB(j) (vB )dvA dvB nA(i) nB(j) (2.15)
ijl all vA all vB

where the integration is over the three velocity components of A and B, respectively.
If we now write the number density of A(i) as

nA(i) = nA pA(i) (2.16)

where nA is the number density of species A and pA(i) is the probability of finding A in
quantum state i, with a similar expression for the B molecules nB(j) = nB pB(j) , a rate
expression equivalent to the well-known phenomenological macroscopic rate expression
is obtained. Thus,

− dnA /dt = kσ nA nB (2.17)

where the rate constant, kσ , is given by


Thermal Equilibrium 29

kσ = pA(i) pB(j) vσR (ij, v|l)fA(i) (vA )fB(j) (vB )dvA dvB
ijl all vA all vB
(2.18)
≡ pA(i) pB(j) kσ (ij|l)
ijl

This equation relates the bimolecular rate constant to the state-to-state rate constant
kσ (ij|l) and ultimately to vσR (ij, v|l). Note that the rate constant is simply the average
value of vσR (ij, v|l). Thus, in a short-hand notation we have kσ = vσR (ij, v|l). The
average is taken over all the microscopic states including the appropriate probability
distributions, which are the velocity distributions fA(i) (vA ) and fB(i) (vB ) in the experiment
and the given distributions over the internal quantum states of the reactants.
Outside high vacuum systems we will have an ensemble of molecules that will
exchange energy. Typically, thermal equilibrium will be maintained during chemical
reaction. There are, though, important exceptions such as chemical reactions in flames
and in explosions, as well as reactions that take place at very low pressures.

2.2 Thermal Equilibrium


We now proceed to develop a specific expression for the rate constant for reactants
where the velocity distributions fA(i) (vA ) and fB(j) (vB ) for the translational motion
are independent of the internal quantum state (i and j) and correspond to thermal
equilibrium.4 Then, according to the kinetic theory of gases or statistical mechanics, see
Appendix B.2.1, Eq. (B.65), the velocity distributions associated with the center-of-mass
motion of molecules are the Maxwell–Boltzmann distribution, a special case of the
general Boltzmann distribution law:

f (vA )dvA = f (vxA , vyA , vzA )dvA


= f (vxA )f (vyA )f (vzA )dvA
 3/2
mA mA v2xA mA v2yA mA v2zA
= exp − exp − exp − dvA
2π kB T 2kB T 2kB T 2kB T
(2.19)
where dvA = dvxA dvyA dvzA , kB is the Boltzmann constant, and a similar expression holds
for molecule B.

4 We assume that the mean-free path is much larger than the molecular dimensions; see Section 9.2. At very
high pressures this “assumption of free flight” is not valid and the overall reaction rate is controlled by the
diffusional motion of the reactants.
30 From Microscopic to Macroscopic Descriptions

Since the relative speed v appears in the integrand in Eq. (2.18), it will be convenient
to change to the center-of-mass velocity V and the relative velocity v (where v = |v|).
We find

v = vA − vB
(2.20)
V = (mA vA + mB vB )/M

where M = mA + mB . The velocities of A and B expressed in terms of the center-of-mass


velocity and the relative velocity are then given by

mB
vA = V + v
M (2.21)
mA
vB = V − v
M

Since the Jacobi determinant for this substitution is equal to one, we have

dvA dvB = dV dv (2.22)

In order to simplify the notation, we consider only the x-component (the derivation is
easily generalized to include all three velocity components). The fraction of A molecules
(mass mA ) with velocity in the range from vxA to vxA + dvxA , at the temperature T , is

mA mA v2xA
f (vxA )dvxA = exp − dvxA (2.23)
2π kB T 2kB T

and the corresponding expression for the B molecules (mass mB ) is

mB mB v2xB
f (vxB )dvxB = exp − dvxB (2.24)
2π kB T 2kB T

We need to evaluate the following product:

f (vxA )f (vxB )dvxA dvxB

appearing in Eq. (2.18). This is the probability of finding an A molecule with velocity in
the range from vxA to vxA + dvxA and a B molecule with velocity in the range from vxB
to vxB + dvxB .
We introduce new variables according to Eq. (2.20):

vx = vxA − vxB
Vx = vxA mA /M + vxB mB /M
Thermal Equilibrium 31

where vx and Vx are the x-component of the relative velocity v and the center-of-
mass velocity V , respectively. The following relation is easily established (see also,
Appendix E.1):

mA v2xA /2 + mB v2xB /2 = μv2x /2 + MVx2 /2

where

μ = mA mB /M (2.25)

is the reduced mass and M is the total mass. The total kinetic energy is, accordingly, equal
to the kinetic energy for the center-of-mass motion plus the kinetic energy of the relative
motion. We now obtain
   
mA mB μv2x MVx2
f (vxA )f (vxB )dvxA dvxB = exp − exp − dvx dVx
(2π kB T )2 2kB T 2kB T

That is, the Maxwell–Boltzmann distribution for the two molecules can be written
as a product of two terms, where the terms are related to the relative motion and the
center-of-mass motion, respectively. After substitution into Eq. (2.18) we √ can perform
the
∞ integration over the
√ center-of-mass velocity V x . This gives the factor 2π kB T /M
( −∞ exp(−ax2 )dx = π/a) and, from the equation before, we obtain the probability
distribution for the relative velocity, irrespective of the center-of-mass motion.
The probability distribution for the relative velocity in the x-direction is accordingly
 1/2  
μ μv2x
f1 (vx )dvx = exp − dvx (2.26)
2π kB T 2kB T

Note that this equation is identical to the Maxwell–Boltzmann velocity distribution of a


single particle with a mass given by the reduced mass.
Due to the simple product form of the Maxwell–Boltzmann distribution, the deriva-
tions given here are easily generalized to the expression for the relative velocity in
three dimensions. Since the integrand in Eq. (2.18) (besides the Maxwell–Boltzmann
distribution) depends only on the relative speed, we can simplify the expression in
Eq. (2.18) further by integrating over the orientation of the relative velocity. This is
done by introducing polar coordinates for the relative velocity. The full three-dimensional
probability distribution for the relative speed is
 3/2  
μ μv2
f3 (v)dv = 4π v2 exp − dv (2.27)
2π kB T 2kB T

where the speed is v = v2x + v2y + v2z , that is, the length of the velocity vector, and the
factor 4π v2 comes from the integration over the orientation of the velocity.
32 From Microscopic to Macroscopic Descriptions

It is useful to determine the corresponding energy distribution, that is, the fraction
of molecules with relative kinetic energy in the range from Etr to Etr + dEtr . We
√ relative kinetic energy is Etr = μv /2, which implies that dEtr = μvdv or
note that the 2

dv = dEtr / 2μEtr . The energy distribution for motion in three dimensions is then
obtained from Eq. (2.27):

 3/2   
1 Etr
f3 (Etr )dEtr = 2π Etr exp − dEtr (2.28)
π kB T kB T

This relation was illustrated in Fig. 1.2.3.


We can now collect the results and obtain an expression for Eq. (2.18) in the case
where the translational motion of the reactants is in thermal equilibrium:


k(T ) = vσR (v)f3 (v)dv
0
∞
= 2Etr /μ σR (Etr )f3 (Etr ) dEtr (2.29)
0
 1/2 ∞  
1 8 Etr
= σR (Etr )Etr exp − dEtr
kB T π μkB T 0 kB T

where

σR (v) = pA(i) pB(j) σR (ij, v|l) (2.30)


ijl

We can also write Eq. (2.29) in the form

 1/2 ∞  
1 8 Etr
k(T ) = pA(i) pB(j) σR (ij, Etr |l)Etr exp − dEtr
kB T π μkB T 0 kB T
ijl
(2.31)
≡ pA(i) pB(j) k(ij|l) (T )
ijl

which defines the state-to-state rate constant k(ij|l) (T ), that is, the rate constant associated
with reactants and products in the specified quantum states (ij|l) with the translational
motion of the reactants in thermal equilibrium.
For reactants in complete internal thermal equilibrium, pA(i) and pB(j) are Boltzmann
distributions (see Appendix B.1), for example,
Thermal Equilibrium 33

pA(i) = exp(−Ei /kB T )/QA (2.32)

is the probability of finding the molecule A in quantum state i with energy Ei and QA is
the molecular partition function,

QA = exp(−Ei /kB T ) (2.33)


i

where the sum includes all internal states of the molecule.

2.2.1 Principle of detailed balance


An additional important general result for k(T ) can be derived in the case of complete
thermal equilibrium, that is, translational as well as complete internal thermal equilib-
rium: the relationship between rate constants for the forward, kf (T ), and the reverse,
kr (T ), reaction

kf
A + B −→
←− C + D
kr

at complete thermal equilibrium is known as the principle of detailed balance,5 and


given by

 3/2  
kf (T ) μCD QC QD
= exp(−E0 /kB T ) ≡ K(T ) (2.34)
kr (T ) μAB QA QB int

where μCD and μAB are the reduced masses of the products and the reactants, respec-
tively, E0 = E0,p − E0,r , E0,p and E0,r are the zero-point energies of the products and
the reactants, respectively, the partition functions refer to the internal (subscript “int” for
internal) motion of the molecules, and K(T ) is the equilibrium constant of the reaction.
This relation is very useful for obtaining information about reverse reactions once the
forward rate constants or cross-sections are known. This principle is a consequence of
microscopic reversibility, that is, the time-reversal symmetry of the equation of motion—
the time-dependent Schrödinger equation (or the classical equations of motion). This
result is proved in Appendix C.

5 This relation between rate constants and the equilibrium constant can, of course, also be derived within
classical chemical kinetics using the equilibrium condition dnA /dt = 0.
34 From Microscopic to Macroscopic Descriptions

Example 2.2 Vibrational relaxation, Boltzmann distribution

This example illustrates several important points.

(i) Inelastic collisions change the internal energy distribution of molecules.


(ii) Inelastic collisions are the physical mechanism behind thermal equilibration, that is, a
Boltzmann distribution is obtained in the long-time limit of vibrational relaxation.
(iii) The relaxation time is inversely proportional to the pressure.

To that end, consider the following inelastic collision:


k10
He + N2 (n = 1) −→
←− He + N2 (n = 0)
k01
where we assume that N2 can only exist in the two vibrational states indicated above. The
bimolecular rate constant of the forward process is k10 , and the rate constant of the reverse
process is k01 (note that the formalism of this chapter also applies to inelastic collisions).
We assume that thermal equilibrium, at the temperature T, has been established in the trans-
lational degrees of freedom. Then, according to the principle of detailed balance (Eq. (C.37)),
k10 /k01 = e−(E0 −E1 )/kB T
where E0 and E1 are the energies associated with the vibrational energy levels. This equation
is similar to Eq. (2.34), except that thermal equilibrium in the vibrational degree of freedom
has not been assumed. The rate constant k10 has been determined at T = 700 K and k10 =
3.0 × 104 liter/(mol s). The initial conditions (t = 0) are [N2 (n = 1)] = [N2 (n = 1)]0 and [N2
(n = 0)] = 0.
The rate law takes the form
d[N2 (n = 1)]
− = k10 [He][N2 (n = 1)] − k01 [He][N2 (n = 0)]
dt
= k10 [He][N2 (n = 1)] − k01 [He]([N2 (n = 1)]0 − [N2 (n = 1)])
= (kf + kr )[N2 (n = 1)] − kr [N2 (n = 1)]0
where kf = k10 [He] and kr = k01 [He]. This is an inhomogeneous first-order differential
equation, with the solution
 
[N2 (n = 1)] = [N2 (n = 1)]0 kr + kf e−(kf +kr )t /(kf + kr )

The concentration in the vibrational ground state is


[N2 (n = 0)] = [N2 (n = 1)]0 − [N2 (n = 1)]
= [N2 (n = 1)]0 (1 − [N2 (n = 1)]/[N2 (n = 1)]0 )
 
= [N2 (n = 1)]0 1 − e−(kf +kr )t kf /(kf + kr )

In the long-time limit, the fraction of molecules in the vibrational ground state becomes
Thermal Equilibrium 35

kf
[N2 (n = 0)]∞ /[N2 (n = 1)]0 =
kf + kr
k10
=
k10 + k01
k10 /k01
=
1 + k10 /k01

e−(E0 −E1 )/kB T


=
1 + e−(E0 −E1 )/kB T

According to the Boltzmann distribution, the probability of observing the N2 molecule in the
vibrational ground state is

e−E0 /kB T
p0 =
Q

e−E0 /kB T
= −E /k T
e 0 B + e−E1 /kB T
e−(E0 −E1 )/kB T
=
1 + e−(E0 −E1 )/kB T

that is, identical to the probability derived from the kinetic approach.
We define the vibrational relaxation time, trelax , as the time it takes to reach 90% of the final
concentration in the vibrational ground state. That is,

[N2 (n = 0)] = 0.9[N2 (n = 0)]∞


kf
= 0.9[N2 (n = 1)]0
kf + kr

which implies that 1 − exp(−(kf + kr )trelax ) = 0.9, and

ln 10
trelax =
[He](k10 + k01 )
RT ln 10
=
pHe k10 (1 + k10 /k01 )

where the ideal gas equation, pHe = [He]RT , was used in the second line. Thus, this equation
shows that the relaxation time is inversely proportional to the pressure. Now for N2 , we have
E0 − E1 = −0.29 eV and trelax = 4.4 × 10−3 s, at pHe = 1 atm. Thus, the vibrational relaxation
is relatively slow, especially at low pressures. Rotational and translation relaxation is much
faster.
36 From Microscopic to Macroscopic Descriptions

Further reading/references
[1] J.C. Light, J. Ross, and K.E. Shuler, in Kinetic processes in gases and plasmas (Academic Press,
1969), Chapter 8.

.....................................................................................................................................

P RO B L E M S

2.1 Using Eq. (2.18), what is the expression for the rate constant if all distributions
associated with internal quantum states and velocities of the reactants are “sharp”?
Mathematically, this condition can be expressed in the following way: pA(i) =
δij and pB(j) = δjk , where the so-called  Kronecker delta is defined by δmn = 1 if
m = n, and δmn = 0 if m = n, that is, m δmn gm = gn . For the continuous velocity
we assume that the distributions are peaked at v0A and v0B , respectively, that is,
fA(i) (vA ) = δ(vA − v0A ) and fB(j) (vB ) = δ(vB − v0B ), where the so-called delta function
   
 ∞ by δ(r − r ) = δ(x − x )δ(y − y )δ(z − z ), where for an arbitrary function
is defined
g(x), −∞ δ(x − x )g(x)dx = g(x ), with equivalent expressions in y and z.
2.2 The equilibrium probability distributions for the relative speed and the relative kinetic
energy are frequently used in various one- or two-dimensional models.
(a) When motion is restricted to two dimensions, show that the probability distri-
bution for the relative speed is
   
μ μv2
f2 (v)dv = 2π v exp − dv
2π kB T 2kB T

where the speed is v = v2x + v2y .
(b) When motion is restricted to one dimension, show that the energy distribution is
 
1 Etr
f1 (Etr )dEtr = √ exp − dEtr
π Etr kB T kB T

Note that the speed is the length of the velocity vector, that is, v = v2x , and
use that the velocity distribution must be multiplied by 2, when the speed is
considered. When motion is restricted to two dimensions, show that the energy
distribution is
 
1 Etr
f2 (Etr )dEtr = exp − dEtr
kB T kB T

2.3 (a) For a reaction without an energy threshold, it was found


√ that the total reaction
cross-section could be represented as σR (Etr ) = A/ Etr , where A is a constant.
Calculate the thermal rate constant k(T ) for the reaction.
Problems 37

In connection with the evaluation of Eq. (2.29), the Gamma function (n)
can sometimes be useful. It is defined as

(n) = xn−1 exp(−x)dx, n > 0
0

with the special values

(n + 1) = n!, where n = 1, 2, . . . ,
1 · 3 · 5 · · · (2n − 1) √
(n + 1/2) = π, where n = 1, 2, . . . .
2n

(b) For a reaction with an energy threshold, it was found that the total reaction
cross-section could be represented as

0 for Etr < E0
σR (Etr ) =
f (Etr − E0 ) for Etr ≥ E0

Show that the thermal rate constant k(T ) for the reaction can be written in
the form k(T ) = A(T )e−E0 /kB T , and specify the connection between A(T ) and
f (Etr − E0 ).

2.4 We consider, in the gas phase, the (SN 2) reaction

Cl− + CH3 Br → ClCH3 + Br−

Reaction cross-sections have been determined in a molecular-beam experiment


[ J. Am. Chem. Soc. 125, 1014 (2003)]. The total cross-section at the relative
translational energy Etr = 0.06 eV is σR (Etr ) = 1.3 × 10−16 cm2 .
(a) Calculate the thermally-averaged rate constant k(T ) at T = 300 K, under the
assumption that in the relevant energy range determined by the Boltzmann
distribution the cross-section is independent of the relative translational energy.
(b) Assume that the differential reaction cross-section, for small deflection
−1/2
angles θ, can be described by the equation dσR /d = AEtr θ −3/2 , where
d = sin θdθdφ and A is a constant. What is the ratio between the number
of product molecules in the two angular regions θ ∈ [0.2, 0.3] radians and
θ ∈ [0.3, 0.366] radians, respectively, when φ ∈ [0.1, 0.2] radians? (Use that, at
small deflection angles sin θ ∼ θ .)
Another random document with
no related content on Scribd:
sauf-conduit, ces présentes vous en serviront.»

—Signé?... fit le malicieux et courageux bossu.


—Signé François, dit Maillé.
—Non, non, reprit le prince, il y a «votre bon cousin et ami
François!»—Messieurs, cria-t-il aux Écossais, je vous suis dans la
prison où vous avez charge de me conduire de la part du roi. Il y a
assez de noblesse en cette salle pour comprendre ceci!
Le profond silence qui régna dans la salle aurait dû éclairer les
Guise; mais le silence est ce que les princes écoutent le moins.
—Monseigneur, dit le cardinal de Tournon qui suivit le prince,
depuis l’affaire d’Amboise, vous avez entrepris sur Lyon et à
Mouvans en Dauphiné des choses contre l’autorité royale,
desquelles le roi n’avait pas connaissance quand il vous écrivait
ainsi.
—Fourbes! s’écria le prince en riant.
—Vous avez fait une déclaration publique contre la messe et
pour l’hérésie...
—Nous sommes maîtres en Navarre, dit le prince.
—Vous voulez dire le Béarn? Mais vous devez hommage à la
couronne, répondit le président de Thou.
—Ah! vous êtes ici, président? s’écria le prince avec ironie. Y
êtes-vous avec tout le parlement?
Sur ce mot, le prince jeta sur le cardinal un regard de mépris et
quitta la salle: il comprit qu’on en voulait à sa tête. Lorsque le
lendemain messieurs de Thou, de Viole, d’Espesse, le procureur-
général Bourdin et le greffier en chef Du Tillet entrèrent dans la
prison, il les tint debout et leur exprima ses regrets de les voir
chargés d’une affaire qui ne les regardait pas; puis il dit au greffier:
Écrivez! et il dicta ceci:

«Moi, Louis de Bourbon, prince de Condé, pair du


royaume, marquis de Conti, comte de Soissons, prince du
sang de France, déclare refuser formellement de reconnaître
aucune commission nommée pour me juger, attendu qu’en
ma qualité et en vertu du privilége attaché à tout membre de
la maison royale, je ne puis être accusé, entendu, jugé, que
par le parlement garni de tous les pairs, toutes les chambres
assemblées, et le roi séant en son lit de justice.»

—Vous deviez savoir cela mieux que d’autres, messieurs, c’est


tout ce que vous aurez de moi. Pour le surplus, je me confie à mon
droit et à Dieu!
Les magistrats procédèrent nonobstant le silence obstiné du
prince. Le roi de Navarre était en liberté, mais observé; sa prison
était plus grande que celle du prince, ce fut toute la différence de sa
position et de celle de son frère; car la tête du prince de Condé et la
sienne devaient tomber du même coup.
Christophe ne fut donc gardé si sévèrement au secret par les
ordres du cardinal et du lieutenant-général du royaume, que pour
donner aux magistrats une preuve de la culpabilité du prince. Les
lettres saisies sur La Sague, le secrétaire du prince, intelligibles pour
des hommes d’État, n’étaient pas assez claires pour des juges. Le
cardinal avait médité de confronter par hasard le prince et
Christophe, qui n’avait pas été placé sans intention dans une salle
basse de la tour de Saint-Agnan, dont la fenêtre donnait sur le
préau. A chaque interrogatoire que les magistrats lui firent subir,
Christophe se renferma dans un système de dénégation absolue,
qui prolongea naturellement le procès jusqu’à l’ouverture des États.
Lecamus, qui n’avait pas manqué de se faire nommer député du
Tiers-État par la bourgeoisie de Paris, arriva quelques jours après
l’arrestation du prince à Orléans. Cette nouvelle, qui lui fut apprise à
Étampes, redoubla ses inquiétudes, car il comprit, lui qui savait seul
l’entrevue du prince et de son fils sous le Pont-au-Change, que le
sort de Christophe était lié à celui de l’audacieux chef du parti de la
Réformation. Aussi résolut-il d’étudier les ténébreux intérêts qui se
croisaient à la cour depuis l’ouverture des États, afin de trouver un
moyen de sauver son fils. Il ne devait pas songer à la reine
Catherine, qui refusa de voir son pelletier. Aucune des personnes de
la cour qu’il put voir ne lui donna de nouvelles satisfaisantes sur son
fils, et il en était arrivé à un tel degré de désespoir, qu’il allait
s’adresser au cardinal lui-même, quand il sut que M. de Thou avait
accepté, ce qui fait une tache à sa vie, d’être un des juges du prince
de Condé. Le syndic alla voir le protecteur de son fils, et apprit que
Christophe était encore vivant, mais prisonnier.
Le gantier Tourillon, chez qui La Renaudie avait envoyé
Christophe, avait offert dans sa maison une chambre au sieur
Lecamus pour tout le temps de la durée des États. Le gantier croyait
le pelletier secrètement attaché, comme lui, à la religion réformée;
mais il vit bientôt qu’un père qui craint pour les jours de son fils ne
comprend plus les nuances religieuses, et se jette à corps perdu
dans le sein de Dieu, sans se soucier de l’écharpe que lui mettent
les hommes. Le vieillard, repoussé dans toutes ses tentatives, allait
comme un hébété par les rues; contre ses prévisions, son or ne lui
servait à rien; monsieur de Thou l’avait prévenu que s’il corrompait
quelque serviteur de la maison de Guise, il en serait pour son
argent, car le duc et le cardinal ne laissaient rien transpirer de ce qui
regardait Christophe. Ce magistrat, dont la gloire est un peu ternie
par le rôle qu’il jouait alors, avait essayé de donner quelque
espérance au père désolé; mais il tremblait tellement lui-même pour
les jours de son filleul, que ses consolations alarmèrent davantage le
pelletier. Le vieillard rôdait autour de la maison. En trois mois, il avait
maigri. Son seul espoir, il le plaçait dans la vive amitié qui depuis
longtemps l’unissait à l’Hippocrate du seizième siècle. Ambroise
essaya de dire un mot à la reine Marie en sortant de la chambre du
roi; mais dès qu’il eut nommé Christophe, la fille des Stuarts, irritée à
la perspective de son sort s’il arrivait malheur au roi, et qui le crut
empoisonné par les Réformés, à cause de l’opportune soudaineté
de sa maladie, répondit:—Si mes oncles m’écoutaient, un pareil
fanatique serait déjà pendu! Le soir où cette funeste réponse fut
donnée à Lecamus par son ami Paré, sur la place de l’Estape, il
revint à demi mort et rentra dans sa chambre en refusant de souper.
Tourillon, inquiet, monta, trouva le vieillard en pleurs, et comme les
yeux vieillis du pauvre pelletier laissaient voir la chair intérieure des
paupières ridées et rougies, le gantier crut qu’il pleurait du sang.
—Consolez-vous, mon père, dit le Réformé, les bourgeois
d’Orléans sont furieux de voir leur ville traitée comme si elle eût été
prise d’assaut, gardée par les soldats de monsieur de Cypierre; et si
la vie du prince de Condé se trouvait en péril, nous aurions bientôt
démoli la tour de Saint-Agnan; car toute notre ville est pour la
Réforme et se révoltera, soyez-en sûr!
—Quand on pendrait les Lorrains, leur mort me rendrait-elle mon
fils? répondit le père désolé.
En ce moment on frappa discrètement à la porte de Tourillon, qui
descendit pour ouvrir lui-même. Il était nuit close. Dans ces temps
de troubles, chaque maître de maison prenait des précautions
minutieuses. Tourillon regarda par la grille du judas pratiqué dans sa
porte, et vit un étranger dont l’accent trahissait un Italien. Cet
homme, vêtu de noir, demandait à parler à Lecamus pour affaires de
commerce, et Tourillon l’introduisit. A la vue de l’étranger, le pelletier
tressaillit horriblement; mais l’étranger trouva le temps de se mettre
un doigt sur les lèvres; Lecamus lui dit alors en comprenant ce
geste: Vous venez sans doute pour m’offrir des fourrures?
—Si, répondit en italien l’étranger d’une façon discrète.
Ce personnage était en effet le fameux Ruggieri, l’astrologue de
la reine-mère. Tourillon descendit chez lui, en comprenant qu’il était
de trop chez son hôte.
—Où pouvons-nous causer sans avoir à craindre qu’on ne nous
entende? dit le prudent Florentin.
—Il nous faudrait être en plein champ, répondit Lecamus; mais
on ne nous laissera pas sortir, vous connaissez la sévérité avec
laquelle les portes sont gardées. Nul ne quitte la ville sans une
passe de monsieur de Cypierre, fût-il, comme moi, membre des
États. Aussi devons-nous dès demain, à notre séance, nous plaindre
tous de ce défaut de liberté.
—Travaillez comme une taupe, mais ne laissez jamais voir vos
pattes dans quoi que ce soit, lui dit le rusé Florentin. La journée de
demain sera sans doute décisive. D’après mes observations, demain
ou après vous aurez peut-être votre fils.
—Que Dieu vous entende, vous qui passez pour ne consulter
que le diable!
—Venez donc chez moi, dit l’astrologue en souriant. J’ai pour
observer les astres la tour du sieur Touchet de Beauvais, le
lieutenant du Bailliage, dont la fille plaît fort au petit duc d’Orléans.
J’ai fait le thème de cette petite, il indique en effet qu’elle sera une
grande dame et aimée par un roi. Le lieutenant est un bel esprit, il
aime les sciences, et la reine m’a fait loger chez ce bonhomme, qui
a l’esprit d’être un forcené guisard en attendant le règne de Charles
IX.
Le pelletier et l’astrologue se rendirent à l’hôtel du sieur de
Beauvais sans être vus ni rencontrés; mais dans le cas où la visite
de Lecamus serait découverte, le Florentin comptait lui donner le
prétexte d’une consultation astrologique sur le sort de Christophe.
Quand ils furent arrivés en haut de la tourelle où l’astrologue avait
mis son cabinet, Lecamus lui dit:—Mon fils est donc bien
certainement vivant?
—Encore, répondit Ruggieri, mais il s’agit de le sauver. Songez,
marchand de peaux, que je ne donnerais pas deux liards de la vôtre,
s’il vous échappait, dans toute votre vie, une seule syllabe de ce que
je vais vous dire.
—Recommandation inutile, mon maître; je suis fournisseur de la
cour depuis le défunt roi Louis XII, et voici le quatrième règne que je
vois.
—Vous direz bientôt le cinquième, repartit Ruggieri.
—Que savez-vous de mon fils?
—Eh! bien, il a été mis à la question.
—Pauvre enfant! dit le bonhomme en levant les yeux au ciel.
—Il a les genoux et les chevilles un tantinet broyés; mais il a
conquis une royale protection qui s’étendra sur toute sa vie, fit
vivement le Florentin en voyant l’effroi du père. Votre petit
Christophe a rendu service à notre grande reine Catherine. Si nous
tirons votre fils des griffes du Lorrain, vous le verrez quelque jour
conseiller au parlement. On se ferait casser trois fois les os pour être
dans les bonnes grâces de cette chère souveraine, un bien beau
génie, qui triomphera de tous les obstacles! J’ai fait le thème du duc
de Guise: il sera tué dans un an d’ici! Voyons, Christophe a vu le
prince de Condé...
—Vous qui savez l’avenir, ne savez-vous point le passé? dit le
pelletier.
—Je ne vous interroge pas, bonhomme, je vous instruis. Or, si
votre fils, qui sera mis demain sur le passage du prince, le reconnaît,
ou si le prince reconnaît votre fils, la tête de monsieur de Condé
sautera. Dieu sait ce qui adviendra de son complice! Rassurez-vous.
Ni votre fils ni le prince ne seront mis à mort, j’ai fait leurs thèmes, ils
doivent vivre; mais j’ignore par quels moyens ils se tireront d’affaire.
Sans compter la certitude de mes calculs, nous allons y mettre
ordre. Demain le prince recevra par des mains sûres un livre de
prières où nous lui ferons passer un avis. Dieu veuille que votre fils
soit discret, car il ne sera pas prévenu, lui! Un seul regard de
connaissance coûtera la vie au prince. Aussi, quoique la reine-mère
ait tout lieu de compter sur la fidélité de Christophe...
—On l’a mise à de rudes épreuves! s’écria le pelletier.
—Ne parlez pas ainsi! Croyez-vous que la reine soit à la noce?
Aussi va-t-elle prendre des mesures comme si les Guise avaient
résolu la mort du prince; et bien fait-elle, la sage et prudente reine!
Or, elle compte sur vous pour être aidée en toute chose. Vous avez
quelque influence sur le Tiers-État, où vous représentez les corps de
métiers de Paris, et quoique les guisards vous promettent de mettre
votre fils en liberté, tâchez de les trupher, et soulevez votre Ordre
contre les Lorrains. Demandez la reine-mère pour régente, le roi de
Navarre y consentira demain publiquement à la séance des États.
—Mais le roi?
—Le roi mourra, répondit Ruggieri, j’ai dressé son thème. Ce que
la reine vous demande de faire pour elle aux États est tout simple;
mais elle attend de vous un plus grand service. Vous avez soutenu
dans ses études le grand Ambroise Paré, vous êtes son ami...
—Ambroise aime aujourd’hui le duc de Guise plus qu’il ne
m’aime, et il a raison, il lui doit sa charge; mais il est fidèle au roi.
Aussi, quoiqu’il incline à la Réforme, ne fera-t-il rien contre son
devoir.
—Peste soit de ces honnêtes gens! s’écria le Florentin. Ambroise
s’est vanté ce soir de tirer le petit roi d’affaire. Si le roi recouvre la
santé, les Guise triomphent, les princes meurent, la maison de
Bourbon sera finie, nous retournerons à Florence, votre fils est
pendu, et les Lorrains auront bon marché des autres enfants de
France...
—Grand Dieu! s’écria Lecamus.
—Ne vous exclamez pas ainsi, c’est d’un bourgeois qui ne sait
rien de la cour; mais allez aussitôt chez Ambroise, et sachez de lui
ce qu’il compte faire pour sauver le roi. S’il y a quelque certitude,
vous viendrez me confier l’opération en laquelle il a tant de foi.
—Mais... dit Lecamus.
—Obéissez aveuglément, mon cher, autrement vous seriez
ébloui.
—Il a raison, pensa le pelletier. Et il alla chez le premier
chirurgien du roi, qui logeait dans une hôtellerie sur la place du
Martroi.
En ce moment, Catherine de Médicis se trouvait dans une
extrémité politique semblable à celle où Christophe l’avait vue à
Blois. Si elle s’était formée à la lutte, si elle avait exercé sa haute
intelligence dans cette première défaite, sa situation, quoique
exactement la même, était aussi devenue plus critique et plus
périlleuse que lors du tumulte d’Amboise. Les événements avaient
grandi autant que la femme. Quoiqu’elle parût marcher d’accord
avec les deux princes lorrains, Catherine tenait les fils d’une
conspiration savamment ourdie contre ses terribles associés, et
attendait un moment propice pour lever le masque. Le Cardinal
venait d’avoir la certitude d’être trompé par Catherine. Cette habile
Italienne avait vu dans la maison cadette un obstacle à opposer aux
prétentions des Guise; et, malgré l’avis des deux Gondi, qui lui
conseillaient de laisser les Guise se porter à des violences contre les
Bourbons, elle avait fait manquer, en avertissant la reine de Navarre,
le projet concerté par les Guise avec l’Espagne de s’emparer du
Béarn. Comme ce secret d’État n’était connu que d’eux et de la
reine-mère, les deux princes lorrains, certains de la duplicité de leur
alliée, voulurent la renvoyer à Florence; et, pour s’assurer de la
trahison de Catherine envers l’État (la maison de Lorraine était
l’État), le duc et le cardinal venaient de lui confier leur dessein de se
défaire du roi de Navarre. Les précautions que prit à l’instant Antoine
de Bourbon prouvèrent aux deux frères que ce secret, connu d’eux
trois seulement, avait été divulgué par la reine-mère. Le cardinal de
Lorraine reprocha sur-le-champ à la reine-mère son manque de foi
devant François II, en la menaçant d’un édit de bannissement, au
cas où de nouvelles indiscrétions mettraient l’État en péril.
Catherine, qui se vit alors dans un extrême danger, devait agir en
grand roi. Aussi donna-t-elle alors la preuve de sa haute capacité;
mais il faut avouer qu’elle fut aussi très-bien servie par ses intimes.
L’Hospital fit parvenir à la Reine un billet ainsi conçu: «Ne laissez
pas mettre à mort un prince du sang par une commission, vous
seriez bientôt enlevée aussi!» Catherine envoya Birague au Vignay,
pour faire dire au chancelier de venir aux États, malgré sa disgrâce.
Birague arriva, cette nuit même, à trois lieues d’Orléans, avec
L’Hospital, qui se déclarait ainsi pour la reine-mère. Chiverny, dont la
fidélité fut alors à bon droit soupçonnée par messieurs de Guise,
s’était sauvé d’Orléans; et, par une marche qui faillit lui coûter la vie,
il avait atteint Écouen en dix heures. Il apprit au connétable de
Montmorency le péril de son neveu, le prince de Condé, et l’audace
des Lorrains. Anne de Montmorency, furieux de savoir que le prince
n’avait dû la vie qu’à la subite invasion du mal dont mourut François
II, arrivait avec quinze cents chevaux et cent gentilshommes. Afin de
mieux surprendre messieurs de Guise, il avait évité Paris en venant
d’Écouen à Corbeil, et de Corbeil à Pithiviers par la vallée de
l’Essonne.
—Capitaine contre capitaine, il y aura peu de laine, dit-il à
l’occasion de cette marche hardie.
Anne de Montmorency, qui avait sauvé la France lors de
l’invasion de Charles-Quint en Provence, et le duc de Guise, qui
avait arrêté la seconde invasion de l’empereur à Metz, étaient en
effet les deux plus grands hommes de guerre de la France à cette
époque. Catherine avait attendu le moment précis de réveiller la
haine du connétable disgracié par les Lorrains. Néanmoins, le
marquis de Simeuse, commandant de Gien, en apprenant l’arrivée
d’un corps aussi considérable que celui mené par le connétable,
sauta sur son cheval, espérant pouvoir prévenir à temps le duc de
Guise. Sûre que le connétable viendrait au secours de son neveu et
pleine de confiance dans le dévouement du chancelier à la cause
royale, la reine-mère avait ranimé les espérances et l’audace du
parti de la Réforme. Les Coligny et les amis de la maison de
Bourbon menacée avaient fait cause commune avec les partisans
de la reine-mère. Une coalition entre des intérêts contraires attaqués
par un ennemi commun, se forma sourdement au sein des États, où
il fut hautement question de nommer Catherine régente du royaume,
dans le cas où François II mourrait. Catherine, dont la foi dans
l’astrologie judiciaire surpassait sa foi en l’Église, avait tout osé
contre ses oppresseurs en voyant son fils mourant à l’expiration du
terme assigné à sa vie par la fameuse sorcière que Nostradamus lui
avait amenée au château de Chaumont.
Quelques jours avant le terrible dénoûment de ce règne,
François II avait voulu se promener sur la Loire, afin de ne pas se
trouver dans la ville au moment où le prince de Condé serait
exécuté. Après avoir abandonné la tête de ce prince au cardinal de
Lorraine, il craignit une sédition tout autant que les supplications de
la princesse de Condé. Au moment de s’embarquer, un de ces vents
frais qui s’élèvent sur la Loire aux approches de l’hiver lui donna un
si cruel mal d’oreille qu’il fut obligé de rentrer; il se mit au lit pour
n’en sortir que mort. En dépit de la controverse des médecins qui,
hormis Chapelain, étaient ses ennemis et ses antagonistes, Paré
soutint qu’un dépôt s’était formé à la tête du roi, et que si l’on ne
donnait pas d’issue aux humeurs, de jour en jour les chances de
mort augmenteraient. Malgré l’heure avancée et la loi du couvre-feu,
sévèrement appliquée dans Orléans, alors exactement en état de
siége, la lampe de Paré brillait à sa croisée, et il étudiait; Lecamus
l’appela d’en bas, et quand il eut crié son nom, le chirurgien ordonna
qu’on ouvrît à son vieil ami.
—Tu ne prends pas de repos, Ambroise, et tout en rendant la vie
aux autres, tu dissiperas la tienne, dit le pelletier en entrant.
Il voyait en effet le chirurgien, ses livres ouverts, ses instruments
épars, devant une tête de mort fraîchement enterré, prise au
cimetière et trouée...
—Il s’agit de sauver le roi...
—En es-tu donc bien certain, Ambroise? s’écria le vieillard en
frémissant.
—Comme de mon existence. Le roi, mon vieux protecteur, a des
humeurs peccantes qui lui pèsent sur le cerveau, qui vont le lui
remplir, et la crise est imminente; mais en lui forant le crâne, je
compte faire sortir ces humeurs et lui dégager la tête. J’ai déjà
pratiqué trois fois cette opération, inventée par un Piémontais, et que
j’ai eu l’heur de perfectionner. La première s’est faite au siége de
Metz, sur monsieur de Pienne, que je tirai d’affaire, et qui depuis
n’en a été que plus sage: il avait un dépôt d’humeurs produit par une
arquebusade au chef. La seconde a sauvé la vie d’un pauvre sur qui
j’eus le désir d’éprouver la bonté de cette audacieuse opération à
laquelle s’était prêté monsieur de Pienne. Enfin, la troisième a eu
lieu à Paris, sur un gentilhomme qui se porte à merveille. Le trépan,
tel est le nom donné à cette invention, est encore peu connu. Les
malades y répugnent, à cause de l’imperfection de l’instrument, que
j’ai fini par améliorer. Je m’essaie donc sur cette tête, afin de ne pas
faillir demain sur celle du roi.
—Tu dois être bien sûr de ton fait, car ta tête serait en danger au
cas où...
—Je gagerais ma vie qu’il sera guéri, répondit Ambroise avec la
sécurité de l’homme de génie. Ah! mon vieil ami, qu’est-ce que
trouer la tête avec précaution? n’est-ce pas faire ce que les soldats
font tous les jours à la guerre sans en prendre aucune?
—Mon enfant, dit l’audacieux bourgeois, sais-tu que sauver le roi,
c’est perdre la France? Sais-tu que cet instrument aura placé la
couronne des Valois sur la tête du Lorrain qui se dit héritier de
Charlemagne? Sais-tu que la chirurgie et la politique sont brouillées
en ce moment? Oui, le triomphe de ton génie est la perte de ta
religion. Si les Guise gardent la régence, le sang des Réformés va
couler à flots! Sois plus grand citoyen que grand chirurgien, et dors
demain la grasse matinée en laissant la chambre libre aux médecins
qui, s’ils ne guérissent pas le roi, guériront la France!
—Moi! s’écria Paré, que je laisse périr un homme quand je puis
le sauver! Non! non, dussé-je être pendu comme fauteur de Calvin,
j’irai de bonne heure à la cour. Ne sais-tu pas que la seule grâce que
je veux demander, après avoir sauvé le roi, est la vie de ton
Christophe. Il y aura certes un moment où la reine Marie ne me
refusera rien.
—Hélas! mon ami, reprit Lecamus, le petit roi n’a-t-il pas refusé la
grâce du prince de Condé à la princesse? Ne tue pas ta religion en
faisant vivre celui qui doit mourir.
—Ne vas-tu pas te mêler de chercher comment Dieu compte
ordonner l’avenir? s’écria Paré. Les honnêtes gens n’ont qu’une
devise: Fais ce que dois, advienne que pourra! Ainsi ai-je fait au
siége de Calais en mettant le pied sur la face du Grand-Maître: je
courais la chance d’être écharpé par tous ses amis, par ses
serviteurs, et je suis aujourd’hui chirurgien du roi; enfin, je suis de la
Réforme, et j’ai messieurs de Guise pour amis. Je sauverai le roi!
s’écria le chirurgien avec le saint enthousiasme de la conviction que
donne le génie, et Dieu sauvera la France.
Un coup fut frappé à la porte, et quelques instants après un
serviteur d’Ambroise remit un papier à Lecamus, qui lut à haute voix
ces sinistres paroles:

«On dresse un échafaud au couvent des Récollets, pour


décapiter demain le prince de Condé.»

Ambroise et Lecamus se regardèrent en proie l’un et l’autre à la


plus profonde horreur.
—Je vais m’en assurer, dit le pelletier.
Sur la place, Ruggieri prit le bras de Lecamus en lui demandant
le secret d’Ambroise pour sauver le roi; mais le vieillard craignit
quelque ruse et voulut aller voir l’échafaud. L’astrologue et le
pelletier allèrent donc de compagnie jusqu’aux Récollets, et
trouvèrent en effet des charpentiers travaillant aux flambeaux.
—Hé! mon ami, dit Lecamus à un charpentier, quelle besogne
faites-vous?
—Nous apprêtons la pendaison des hérétiques, puisque la
saignée d’Amboise ne les a pas guéris, dit un jeune Récollet qui
surveillait les ouvriers.
—Monseigneur le cardinal a bien raison, dit le prudent Ruggieri;
mais dans notre pays, nous faisons mieux.
—Et que faites-vous? dit le Récollet.
—Mon frère, on les brûle.
Lecamus fut obligé de s’appuyer sur l’astrologue, ses jambes
refusaient de le porter; car il pensait que son fils pouvait demain être
accroché à l’une de ces potences. Le pauvre vieillard était entre
deux sciences, entre l’astrologie judiciaire et la chirurgie, qui toutes
deux lui promettaient le salut de son fils pour qui l’échafaud se
dressait évidemment. Dans le trouble de ses idées, il se laissa
manier comme une pâte par le Florentin.
—Eh! bien, mon respectable marchand de menu-vair, que dites-
vous de ces plaisanteries lorraines? fit Ruggieri.
—Hélas! vous savez que je donnerais ma peau pour voir saine et
sauve celle de mon fils!
—Voilà qui est parler en marchand d’hermine, reprit l’Italien; mais
expliquez-moi bien l’opération que compte faire Ambroise sur le roi,
je vous garantis la vie de votre fils...
—Vrai! s’écria le vieux pelletier.
—Que voulez-vous que je vous jure?... fit Ruggieri.
Sur ce mouvement, le pauvre vieillard répéta son entretien avec
Ambroise au Florentin qui laissa dans la rue le père au désespoir,
dès que le secret du grand chirurgien lui fut divulgué.
—A qui diable en veut-il, ce mécréant! s’écria le vieillard en
voyant Ruggieri se dirigeant au pas de course vers la place de
l’Estape.
Lecamus ignorait la scène terrible qui se passait autour du lit
royal, et qui avait motivé l’ordre d’élever l’échafaud du prince dont la
condamnation avait été prononcée par défaut, pour ainsi dire, et
dont l’exécution avait été remise à cause de la maladie du roi.
Il ne se trouvait dans la salle, dans les escaliers et dans la cour
du Bailliage, que les gens absolument de service. La foule des
courtisans encombrait l’hôtel du roi de Navarre, à qui la régence
appartenait d’après les lois du royaume. La noblesse française,
effrayée d’ailleurs par l’audace des Guise, éprouvait le besoin de se
serrer autour du chef de la maison cadette, en voyant la reine-mère
esclave des Guise et ne comprenant pas sa politique d’Italienne.
Antoine de Bourbon, fidèle à son accord secret avec Catherine, ne
devait renoncer en sa faveur à la régence qu’au moment où les
États prononceraient sur cette question. Cette solitude profonde
avait agi sur le Grand-Maître, quand, au retour d’une ronde faite par
prudence dans la ville, il ne trouva chez le roi que les amis attachés
à sa fortune. La chambre où l’on avait dressé le lit de François II est
contiguë à la grande salle du Bailliage. Elle était alors revêtue de
boiseries en chêne. Le plafond, composé de petites planches
longues savamment ajustées et peintes, offrait des arabesques
bleues sur un fond d’or, dont une partie arrachée il y a cinquante ans
bientôt a été recueillie par un amateur d’antiquités. Cette chambre
tendue de tapisseries et sur le plancher de laquelle s’étendait un
tapis, était si sombre, que les torchères allumées y jetaient peu de
lumière. Le vaste lit, à quatre colonnes et à rideaux de soie,
ressemblait à un tombeau. D’un côté de ce lit, au chevet, se tenaient
la reine Marie et le cardinal de Lorraine. Catherine était assise dans
un fauteuil. Le fameux Jean Chapelain, médecin de service, et qui
fut depuis le premier médecin de Charles IX, se trouvait debout à la
cheminée. Le plus grand silence régnait. Le jeune roi, maigre, pâle,
comme perdu dans ses draps, laissait à peine voir sur l’oreiller sa
petite figure grimée. La duchesse de Guise, assise sur une
escabelle, assistait la jeune reine Marie, et du côté de Catherine,
dans l’embrasure de la croisée, madame de Fiesque épiait les
gestes et les regards de la reine-mère, car elle connaissait les
dangers de sa position.
Dans la salle, malgré l’heure avancée de la soirée, monsieur de
Cypierre, gouverneur du duc d’Orléans, et nommé gouverneur de la
ville, occupait un coin de la cheminée avec les deux Gondi. Le
cardinal de Tournon, qui dans cette crise épousa les intérêts de la
reine-mère en se voyant traité comme un inférieur par le cardinal de
Lorraine, de qui certes il était ecclésiastiquement l’égal, causait à
voix basse avec les Gondi. Les maréchaux de Vieilleville et de Saint-
André, le garde-des-sceaux, qui présidait les États, s’entretenaient à
voix basse des dangers auxquels les Guise étaient exposés.
Le lieutenant-général du royaume traversa la salle en y jetant un
rapide coup d’œil, et y salua le duc d’Orléans qu’il y aperçut.
—Monseigneur, dit-il, voici qui peut vous apprendre à connaître
les hommes: la noblesse catholique du royaume est chez un prince
hérétique, en croyant que les États donneront la régence aux
héritiers du traître qui fit retenir si longtemps en prison votre illustre
grand-père!
Puis, après ces paroles destinées à faire un profond sillon au
cœur d’un prince, il passa dans la chambre, où le jeune roi était
alors moins endormi que plongé dans une lourde somnolence.
Ordinairement, le duc de Guise savait vaincre par un air très-affable
l’aspect sinistre de sa figure cicatrisée; mais en ce moment il n’eut
pas la force de sourire en voyant se briser l’instrument de son
pouvoir. Le cardinal, qui avait autant de courage civil que son frère
avait de courage militaire, fit deux pas et vint à la rencontre du
lieutenant-général.
—Robertet croit que le petit Pinard est vendu à la reine-mère, lui
dit-il à l’oreille en l’emmenant dans la salle, on s’est servi de lui pour
travailler les membres des États.
—Eh! qu’importe que nous soyons trahis par un secrétaire quand
tout nous trahit! s’écria le lieutenant-général. La ville est pour la
Réformation, et nous sommes à la veille d’une révolte. Oui! les
Guépins sont mécontents, reprit-il en donnant aux Orléanais leur
surnom, et si Paré ne sauve pas le roi, nous aurons une terrible
levée de boucliers. Avant peu de temps nous aurons à faire le siége
d’Orléans qui est une crapaudière de Huguenots.
—Depuis un moment, reprit le cardinal, je regarde cette Italienne
qui reste là dans une insensibilité profonde, elle guette la mort de
son fils, Dieu lui pardonne! je me demande si nous ne ferions pas
bien de l’arrêter, ainsi que le roi de Navarre.
—C’est déjà trop d’avoir en prison le prince de Condé! répondit le
duc.
Le bruit d’un cavalier arrivant à bride abattue retentit à la porte du
Bailliage. Les deux princes lorrains allèrent à la fenêtre, et à la lueur
des torches du concierge et de la sentinelle qui brûlaient toujours
sous le porche, le duc reconnut au chapeau cette fameuse croix de
Lorraine que le cardinal venait de faire prendre à ses partisans. Il
envoya l’un des arquebusiers, qui étaient dans l’antichambre, dire de
laisser entrer le survenant, à la rencontre duquel il alla sur le palier,
suivi de son frère.
—Qu’y a-t-il, mon cher Simeuse? demanda le duc avec le
charme de manières qu’il déployait pour les gens de guerre en
voyant le gouverneur de Gien.
—Le connétable entre à Pithiviers, il a quitté Écouen avec quinze
cents chevaux d’ordonnance et cent gentilshommes...
—Sont-ils accompagnés? dit le duc.
—Oui, monseigneur, répondit Simeuse, ils sont en tout deux mille
six cents. Thoré, selon quelques-uns, est en arrière avec un parti
d’infanterie. Si le connétable s’amuse à attendre son fils, vous avez
le temps de le défaire...
—Vous ne savez rien de plus? Les motifs de cette prise d’armes
sont-ils répandus?
—Anne parle aussi peu qu’il écrit, allez à sa rencontre, mon frère,
pendant que je vais le saluer avec la tête de son neveu, dit le
cardinal en donnant l’ordre d’aller chercher Robertet.
—Vieilleville! cria le duc au maréchal qui vint, le connétable a
l’audace de se présenter en armes, si je vais à sa rencontre,
répondez-vous de maintenir la ville?
—Dès que vous sortirez, les bourgeois prendront les armes. Et
qui peut savoir le résultat d’une affaire entre des cavaliers et des
bourgeois au milieu de ces rues étroites? répondit le maréchal.
—Monseigneur, dit Robertet en montant précipitamment
l’escalier, le chancelier est aux portes et veut entrer, doit-on lui
ouvrir?
—Ouvrez, répondit le cardinal de Lorraine. Connétable et
chancelier ensemble, ils seraient trop dangereux, il faut les séparer.
Nous avons été rudement joués par la reine-mère dans le choix de
L’Hospital pour cette charge.
Robertet fit un signe de tête à un capitaine qui attendait une
réponse au bas de l’escalier, et se retourna vivement pour écouter
les ordres du cardinal.
—Monseigneur, je prends la liberté, dit-il en faisant encore un
effort, de représenter que la sentence doit être approuvée par le roi
en son conseil. Si vous violez la loi pour un prince du sang, on ne la
respectera ni pour un cardinal, ni pour un duc de Guise.
—Pinard t’a dérangé, Robertet, dit sévèrement le cardinal. Ne
sais-tu pas que le roi a signé l’arrêt, le jour où il est sorti pour nous le
laisser exécuter!
—Quoique vous me demandiez à peu près ma tête en me
commettant à cet office, qui sera d’ailleurs exécuté par le prévôt de
la ville, j’y vais, monseigneur.
Le Grand-Maître entendit ce débat sans sourciller; mais il prit son
frère par le bras et l’emmena dans un coin de la salle.
—Certes, lui dit-il, les héritiers de Charlemagne ont le droit de
reprendre une couronne qui fut usurpée par Hugues Capet sur leur
maison; mais le peuvent-ils? La poire n’est pas mûre. Notre neveu
se meurt, et toute la cour est chez le roi de Navarre.
—Le cœur a failli au roi. Sans cela, le Béarnais eût été dagué,
reprit le cardinal, et nous aurions eu bon marché de tous les enfants.
—Nous sommes mal placés ici, dit le duc. La sédition de la ville
serait appuyée par les États. L’Hospital, que nous avons tant
protégé, et à l’élévation duquel a résisté la reine Catherine, est
aujourd’hui contre nous, et nous avons besoin de la justice. La reine-
mère est soutenue par trop de monde aujourd’hui, pour que nous
puissions la renvoyer... D’ailleurs, encore trois princes!
—Elle n’est plus mère, elle est toute reine, dit le cardinal; aussi,
selon moi, serait-ce le moment d’en finir avec elle. De l’énergie et
encore de l’énergie! voilà mon ordonnance.
Après ce mot, le cardinal rentra dans la chambre du roi, suivi du
Grand-Maître. Ce prêtre alla droit à Catherine.
—Les papiers de La Sague, secrétaire du prince de Condé, vous
ont été communiqués, vous savez que les Bourbons veulent
détrôner vos enfants? lui dit il.
—Je sais tout cela, répondit l’Italienne.
—Hé! bien, voulez-vous faire arrêter le roi de Navarre?
—Il y a, dit-elle, un lieutenant-général du royaume.
En ce moment, François II se plaignit de douleurs violentes à
l’oreille et se mit à geindre d’un ton lamentable. Le médecin quitta la
cheminée où il se chauffait et vint examiner l’état de la tête.
—Hé! bien, monsieur? dit le Grand-Maître en s’adressant au
premier médecin.
—Je n’ose prendre sur moi d’appliquer un cataplasme pour
attirer les humeurs. Maître Ambroise a promis de sauver le roi par
une opération, je la contrarierais.
—Remettons à demain, dit froidement Catherine, et que tous les
médecins y soient, car vous savez les calomnies auxquelles donne
lieu la mort des princes.
Elle alla baiser la main de son fils et se retira.
—Avec quelle tranquillité cette audacieuse fille de marchand
parle de la mort du dauphin empoisonné par Montecuculli, un
Florentin de sa suite! s’écria la reine Marie Stuart.
—Marie! cria le petit roi, mon grand-père n’a jamais mis son
innocence en doute!...
—Peut-on empêcher cette femme de venir demain? dit la reine à
ses deux oncles à voix basse.
—Que deviendrions-nous, si le roi mourait? répondit le cardinal,
Catherine nous ferait rouler tous dans sa tombe.
Ainsi la question fut nettement posée pendant cette nuit entre
Catherine de Médicis et la maison de Lorraine. L’arrivée du
chancelier et celle du connétable indiquaient une révolte, la matinée
du lendemain allait donc être décisive.
Le lendemain, la reine-mère arriva la première. Elle ne trouva
dans la chambre de son fils que la reine Marie Stuart, pâle et
fatiguée, qui avait passé la nuit en prières auprès du lit. La duchesse
de Guise avait tenu compagnie à la reine, et les filles d’honneur
s’étaient relevées. Le jeune roi dormait. Ni le duc, ni le cardinal
n’avaient encore paru. Le prêtre, plus hardi que le soldat, déploya,
dit-on, dans cette dernière nuit, toute son énergie, sans pouvoir
décider le duc à se faire roi. En face des États-Généraux assemblés,
et menacé d’une bataille à livrer au connétable de Montmorency, le
Balafré ne trouva pas les circonstances favorables; il refusa d’arrêter
le roi de Navarre, la reine-mère, le chancelier, le cardinal de
Tournon, les Gondi, Ruggieri et Birague, en objectant le soulèvement
qui suivrait des mesures si violentes. Il subordonna les projets de
son frère à la vie de François II.
Le plus profond silence régnait dans la chambre du roi.
Catherine, accompagnée de madame de Fiesque, vint au bord du lit
et contempla son fils d’un air dolent admirablement joué. Elle se mit
son mouchoir sur les yeux et alla dans l’embrasure de la croisée, où
madame de Fiesque lui apporta un siége. De là, ses yeux
plongeaient sur la cour.
Il avait été convenu entre Catherine et le cardinal de Tournon,
que si le connétable entrait heureusement en ville, le cardinal
viendrait accompagné des deux Gondi, et qu’en cas de malheur, il
serait seul. A neuf heures du matin, les deux princes lorrains, suivis
de leurs gentilshommes qui restèrent dans le salon, se montrèrent
chez le roi; le capitaine de service les avait avertis qu’Ambroise Paré
venait d’y arriver avec Chapelain et trois autres médecins suscités
par Catherine, qui tous trois haïssaient Ambroise.
Dans quelques instants, la grande salle du Bailliage offrit
absolument le même aspect que la salle des gardes à Blois, le jour
où le duc de Guise fut nommé lieutenant-général du royaume, et où
Christophe fut mis à la torture, à cette différence près, qu’alors
l’amour et la joie remplissaient la chambre royale, que les Guise
triomphaient; tandis que le deuil et la mort y régnaient, et que les
Lorrains sentaient le pouvoir leur glisser des mains. Les filles des
deux reines étaient en deux camps à chaque coin de la grande
cheminée, où brillait un énorme feu. La salle était pleine de
courtisans. La nouvelle répandue, on ne sait par qui, d’une
audacieuse conception d’Ambroise pour sauver les jours du roi,
amenait tous les seigneurs qui avaient droit d’entrer à la cour.
L’escalier extérieur du Bailliage et la cour étaient pleins de groupes
inquiets. L’échafaud dressé pour le prince en face du couvent des
Récollets étonnait toute la noblesse. On causait à voix basse, et les
discours offraient, comme à Blois, le même mélange de propos
sérieux, frivoles, légers et graves. On commençait à prendre
l’habitude des troubles, des brusques révolutions, des prises
d’armes, des rébellions, des grands événements subits qui
marquèrent la longue période pendant laquelle la maison de Valois
s’éteignit, malgré les efforts de la reine Catherine. Il régnait un
profond silence à une certaine distance autour de la porte de la
chambre du roi, gardée par deux hallebardiers, par deux pages et
par le capitaine de la garde écossaise. Antoine de Bourbon,
emprisonné dans son hôtel, y apprit, en s’y voyant seul, les
espérances de la cour, et fut accablé par la nouvelle des apprêts
faits pendant la nuit pour l’exécution de son frère.
Devant la cheminée du Bailliage était l’une des plus belles et plus
grandes figures de ce temps, le chancelier de L’Hospital, dans sa
simarre rouge à retroussis d’hermine, couvert de son mortier, suivant
le privilége de sa charge. Cet homme courageux, en voyant des
factieux dans ses bienfaiteurs, avait épousé les intérêts de ses rois,
représentés par la reine-mère; et, au risque de perdre la tête, il était
allé se consulter avec le connétable, à Écouen; personne n’osait le
tirer de la méditation où il était plongé. Robertet, le secrétaire d’État,
deux maréchaux de France, Vieilleville et Saint-André, le garde-des-
sceaux, formaient un groupe devant le chancelier. Les courtisans ne
riaient pas précisément; mais leurs discours étaient malicieux, et
surtout chez ceux qui ne tenaient pas pour les Guise.
Le cardinal avait enfin saisi l’Écossais Stuart, l’assassin du
président Minard, et faisait commencer son procès à Tours. Il gardait
également, dans le château de Blois et dans celui de Tours, un
assez bon nombre de gentilshommes compromis, pour inspirer une
sorte de terreur à la noblesse, qui ne se terrifiait point, et qui
retrouvait dans la Réformation un appui pour cet amour de révolte
inspiré par le sentiment de son égalité primitive avec le roi. Or, les
prisonniers de Blois avaient trouvé moyen de s’évader, et, par une
singulière fatalité, les prisonniers de Tours venaient d’imiter ceux de
Blois.
—Madame, dit le cardinal de Châtillon à madame de Fiesque, si
quelqu’un s’intéresse aux prisonniers de Tours, ils sont en grand
danger.
En entendant cette phrase, le chancelier tourna la tête vers le
groupe des filles de la reine-mère.
—Oui, le jeune Desvaux, l’écuyer du prince de Condé, qu’on
retenait à Tours, vient d’ajouter une amère plaisanterie à sa fuite. Il
a, dit-on, écrit à messieurs de Guise ce petit mot: «Nous avons
appris l’évasion de vos prisonniers de Blois; nous en avons été si
fâchés, que nous nous sommes mis à courir après eux; nous vous
les ramènerons dès que nous les aurons arrêtés.»
Quoique la plaisanterie lui allât, le chancelier regarda monsieur
de Châtillon d’un air sévère. On entendit en ce moment des voix

You might also like