Вы находитесь на странице: 1из 63

Chapter 4: Chemical Reaction Dynamics

a)

Chemical reaction dynamics is concerned with


unraveling the mechanism of chemical reactions on a
quantum mechanical level. Some key questions:

How does the BO-PES influence a chemical


reaction ? What are the driving forces behind a b)
chemical process ?
How does the kinetic energy and the internal
quantum state of the reactants (electronic,
vibrational, rotational) influence the chemical
reactivity ?
Which reaction product channels are available and
how is energy partitioned between them ?
What are the physical constraints on a chemical
reaction, i.e., are there chemical “selection
rules” ? What is the role of angular momentum ?
Chapter 4: Chemical Reaction Dynamics
a)

Recommended literature:

• M. Brouard, C. Vallance (eds.), Tutorials in


Molecular Reaction Dynamics, RSC Publishing 2010
• R.D. Levine, Molecular Reaction Dynamics,
Cambridge University Press 2005 b)
• M. Brouard, Reaction Dynamics, Oxford Chemistry
Primers, Oxford University Press 1998
• H.H. Telle, A.G. Urena, R.J. Donovan, Laser
Chemistry, Wiley 2007
• B.J. Whitaker, Imaging in Molecular Dynamics,
Cambridge University Press 2003
• M.S. Child, Molecular Collision Theory, Dover 1996
• J.Z. Zhang, Theory and Application of Quantum
Molecular Dynamics, World Scientific 1999
a)
Chapter 4: Contents

4.1 Reaction rates and cross sections


4.2 Classical scattering theory b)
4.3 Introduction to quantum scattering theory
4.4 Reactive scattering: concepts, methods and examples
4.5 Photodissociation dynamics and laser chemistry
4.6 Real-time studies of reactions: Femtochemistry
4.1 Reaction rates and cross sections
4.1.1 Rate constants

Pro memoria: the molecularity is defined as the number of particles involved in


an elementary chemical reaction:
unimolecular: A→B
bimolecular: A+B→C
The rate v for a bimolecular reaction is given by
dCA
v =− = k(T )CA CB
dt
thermal number
rate coefficient densities

If [C]=[molecules cm-3], then the dimension of the rate constant k(T) is [k]=[cm3
molec.-1 s-1] or simply [k]=[cm3 s-1].

In many cases, the temperature dependence of the thermal rate coefficient can
be described in terms of the empirical Arrhenius equation:
k(T ) = Ae −EA /RT
pre-exponential
activation energy
factor
4.1.2 Reaction cross sections

Consider an experiment in which a beam of molecules A with intensity I0 enters


a chamber filled with a gas of molecules B. A reacts with B, and after passing a
distance l through the chamber, the intensity is reduced from I0 to I1 because of
reactive collisions.

A
I0
B I1

The intensity I of the beam of molecules A (molecules passing through a


surface per second) is given by
I = vA CA
velocity number density
If we assume that the B molecules are much slower than the molecular beam of
A molecules (vB=0), the attenuation of the intensity of the beam can be cast into
a Lambert-Beer-type form of expression:
dI
− = σCB I (*)
dx
reaction cross
section

Integrate: ln(I1 /I0 ) = σCB �


The bimolecular rate constant for the reaction is defined as:
dCA
− = kCA CB (**)
dt
Using I=vACA and vA=dx/dt 1/dx=vA/dt, the left-hand side of Eq. (*) becomes:
dI d(CA vA ) dCA dCA
− =− = −vA =−
dx dx dx dt
Using I=vACA , the right-hand side of Eq. (*) becomes:
σCB I = σvA CA CB
dCA
Thus, Eq. (*) becomes: − = σvA CA CB
dt
Comparison with (**) yields: k = σvA

which is an universal expression linking the rate constant with the cross section.
Thermal averaging: In a molecular beam, the molecules have a well-defined
kinetic energy E, thus defining a rate constant k(E). The thermal rate constant
k(T) is obtained by averaging the over the Maxwell-Boltzmann distribution p(E)
of all kinetic energies E: � ∞
�k(E)� ≡ k(T ) = p(E)k(E)dE
0
�√ �−1 � �
√ π E ... Maxwell-Boltzmann
where p(E) = E (kB T )3/2 exp −
2 kB T kinetic-energy distribution
Inserting p(E) and using k(E)=σv=σ(2E/μ)1/2 where μ is the reduced mass we get:
� ∞ √ �√ �−1 � �

π 3/2 E
�k(E)� = E (kB T ) exp − σ(E) 2E/µ dE
0 2 kB T
Changing to the dimensionless energy variable (E/kBT) leads to:
� �1/2 � ∞ � � � �� �
8kB T E E dE
�k(E)� = σ(E) exp −
πµ 0 kB T kB T kB T
average thermal velocity〈vrel〉

This expression can be formulated as


k(T ) = �vrel � · �σ�
where � ∞� � � �� �
E E dE
�σ� = σ(E) exp − ... averaged cross section
0 k B T k B T k B T
The classical collision density ZA (defined as the number of collision per
second) of a molecule A with molecules B is given by:
ZA = k(T )CAB = �σ��vrel �CB
The number of collisions between A and B molecules per unit volume is thus:
V
ZAB = �σ��vrel �CA CB

If A=B, we get: V
ZAA = (1/2)�σ��vrel �CA2 (Factor 1/2 for not counting collisions between the same
particles twice)

4.1.3 Simple collision models for the reaction cross section


1. Hard-sphere collisions: constant reaction cross section σ0
Molecules are treated as colliding hard spheres with radius rA and
rB. Assuming every collision leads to reaction, the cross section is
thus given by: σ0 = πd 2 = π(rA + rB )2
The thermal rate constant is:
� �1/2 � ∞ � � � �� �
8kB T E E dE
k(T ) = σ0 exp −
πµ 0 kB T kB T kB T
Evaluation of the integral
� conveniently
�1/2 yields 1:
8kB T
k(T ) = σ0
πµ
2. The impact parameter b:
The impact parameter b is defined as the distance of
closest approach of the reactants in the absence of an
interaction potential:
• b≈0: head-on collision
• b>>0: glancing collision
The reaction cross section can generally be formulated as
� bmax
σ= P (b)2πb db
0
Where P(b) is the probability for reaction at collision at a given value of b (the
opacity function). b, P(b) and σ are usually dependent on the collision energy.
If P(b)=1, we recover the hard-sphere collision model:
� bmax
2
σ= 2πb db = πbmax
0

3. Reactions with a threshold (activation) energy E0:


Reaction only occurs if E>E0:
� �1/2 � ∞ � � � �� �
8kB T E E dE
k(T ) = σ0 exp −
πµ E0 kB T kB T kB T
Integration yields:
� �1/2 � � � �
8kB T E0 E0
k(T ) = σ0 1+ exp −
πµ kB T kB T

which is of the same form as the Arrhenius equation


k(T ) = Ae −EA /RT
if EA is identified with E0 and A is interpreted as a term A(T) which varies only
slowly with temperature:
� �1/2 � �
8kB T E0
A(T ) = σ0 1+
πµ kB T

In this way, the Arrhenius equation can be derived within the framework of simple
classical collision theory.
4.2 Classical scattering theory

Every chemical reaction entails a collision, a scattering event. We will therefore


treat chemical reactions in the framework of scattering theory.

Molecules are quantum systems - so why use classical models ?


• The essential physical concepts are much easier to understand in a classical
picture
• Classical scattering models are still used for even rather small molecules (>3
atoms !) for which a quantum treatment is prohibitively expensive

Types of scattering events:


• Elastic scattering: total kinetic energy and the internal state of the collision
partners are conserved
• Inelastic scattering: total kinetic energy and internal state of the reaction
partners change, the chemical structure is conserved
• Reactive scattering: kinetic energy, internal state and chemical structure
change
4.2.1 Kinematics of molecular collisions:
the centre-of-mass system

For collisions between molecules, the relevant kinematics are defined by their
motion relative to one another, and not by their absolute motion in the laboratory
coordinate frame.

On thus transforms the system into the centre-of-mass A R


coordinate frame defined by y c.o.m.
rA Rc B
mA�rA + mB �rB

Rc = ... coordinate vector of the
mA + mB rB
centre of mass (c.o.m.)
x
� = �rA − �rB
R ... relative coordinate vector

It can easily be shown that the kinetic energy Ekin of the system is given by:
1 2 1 2 1 2 1 2
Ekin = mA vA + mB vB = MV + µv
2 2 2 2
�˙
�˙ c | , v = |R|
where the velocities are given by vi = |�r˙i |,V = |R
mA mB
M = mA + mB is the total mass and µ = is the reduced mass.
mA + mB
4.2.2 Elastic scattering

Elastic collisions, i.e., collisions in which the kinetic energy is conserved, are
the simplest form of scattering events.

We will discuss classical elastic collisions to introduce the basic concepts of


scattering theory.

Consider the collision trajectory of two structureless particles


(e.g., atoms) in the COM frame:

As the total angular momentum is

y
or
ct
je
conserved, two coordinates suffice to

tra
on
describe the relative motion of the collision

si
lli
co
partners. We choose initial
velocity vector deflection angle
R ... relative position vector
ψ ... orientation angle of R with
respect to the original velocity vector of
closest approach
vector v
relative orientation
impact
position vector angle
parameter
Conserved physical quantities

� = µ�
Angular momentum: L �
v ×R

L before collision = L after collision:

y
or
ct
je
tra
� = L = |µ�
|L| � where v is the initial velocity vector
v × R|

on
si
lli
� sin Ψ
v ·R
= µ�

co
initial
velocity vector deflection angle
� = L = µv b
|L|

vector of
Total energy: closest approach

relative orientation
E = Ekin + Ecent + Epot impact
position vector angle
parameter
kinetic centrifugal potential

1 1
= µṘ + µR2 Ψ̇2 + V (R) with the angular velocity Ψ̇ = dΨ/dt = ω
2
2 2
2 2
1 1 L 1 1 L
= µṘ2 + 2
+ V (R) ⇒ E = µṘ2 + 2
+ V (R)
2 2 µR 2 2 µR
L = µR2 ω
VL(R) ... centrifugally corrected (effective) potential
Centrifugally corrected potentials

Centrifugal energy = energy taken up in the


rotation of the position vector R
Collisional angular momentum L = angular
momentum associated with the rotation of R
about ψ
The effective potential for the collision contains
both, the interaction potential V(R) and the
centrifugal energy:
centrifugal barrier
1 L2
VL (R) = 2
+ V (R)
2 µR
Centrifugally corrected potentials
VL(R) for L3 > L2 > L1 > L0=0
The deflection function χ(b)

The angle of deflection χ depends on the impact parameter b.

Examples:

1. Hard-sphere collisions (a billiard game):

• For b > d: χ = 0
• For b < d: χ = π − 2ψ0 where b/d = sin ψ0 =⇒ χ = 2 arccos(b/d)
2. General potentials with repulsive and attractive parts:

repulsive, short range


part: V(R)>0

χg
b* = b/Re

V(R)
R*

attractive, long range


part: V(R)<0

χr
• Small b: collision dominated by repulsive forces backward scattering
• Large b: collision dominated by attractive forces forward scattering
• Rainbow angle χr: maximum negative deflection angle at impact parameter
br≈Re where the potential is most attractive
• Glory angle χg: deflection angle at impact parameter bg≈R* where attractive
and repulsive forces cancel
• Experimentally, it is not possible to distinguish between positive and negative
deflection angles χ because of the cylindrical symmetry of the collision process.
One can only measure the absolute value of the deflection angle θ=|χ|.
Experimental observables in molecular-collision experiments

The intensity of scattered molecules I(Ω), i.e., the flux of molecules scattered
into the solid angle Ω, defines the differential cross section dσ/dΩ:

dσ scattered flux of molecules per unit solid angle
I(Ω) = = θ
dΩ incident flux of molecules per unit area

The integral cross section σ is obtained by integration.


� � π
dσ dσ
σ= dΩ = 2π sin θdθ
dΩ 0 dΩ
Where the cylindrical symmetry of the problem allowed us to express
dΩ = 2π sin θ dθ
in the second step.

The scattering rate constant is then given by (see section 4.1.2):

k = σv
Calculating the differential cross section from the deflection function θ(b)

If we assume that the opacity function is unity, P(b)=1, the differential cross
section can be expressed as (see sec. 4.1.3)
dσ = 2πb db
Again, because the scattering problem is cylindrically symmetric, the solid angle
element dΩ can be formulated as
dΩ = 2π sin θ dθ
Hence we obtain for the differential cross section:
dσ 2πb db b
= I(θ) = =
dΩ 2π sin θ dθ sin θ(dθ/db)
If more than one value of b contribute to the same scattering angle θ, we have
to sum over all contributions and arrive at the following dependence of the
differential cross section on the deflection function θ(b):
dσ � b
= I(θ) =
dΩ sin θ(dθ/db)

Singularities in the differential cross section (dσ/dΩ=∞):


• Glory (θ=0) singularity: sin θ = 0
• Rainbow singularity: (dθ/db) = 0 (maximum of the function θ(b) )
Illustration:

dσ � b deflection function θ(b)


= I(θ) =
dΩ sin θ(dθ/db)

differential cross
section I(θ)

glory rainbow
singularity singularity
Calculating the deflection function θ(b) from the potential V(R)
It can be shown (see, e.g., R.D. Levine, Molecular Reaction Dynamics):
� ∞
b 1
χ(b) = π − � �1/2 dR
−∞ R2 b2 V (R)
1− R2 − E

i.e., χ(b) depends on the potential V(R) and the collision energy E.
For inverse power law potentials
Cn
V (R) = n
R
which describe long-range interactions between molecules the deflection function
can be approximated to:
V (b)
χ(b) ≈
E

in the limit of large impact parameters b (momentum approximation). Hence, in this


limit the deflection function is a direct measure of the potential !

In an experiment, the impact parameter b cannot be selected and one measures


a differential cross section summed over all possible impact parameters.
4.3 Introduction to quantum scattering theory

4.3.1 Quantum elastic scattering


Contents:
4.3.1.1 General formulation of the scattering problem
4.3.1.2 The scattering phase
4.3.1.3 Scattering amplitude and scattering matrix
Derivation → blackboard

4.3.2 Quantum inelastic scattering


Contents:
4.3.2.1 Scattering Hamiltonian
4.3.2.2 Angular momenta
4.3.2.3 Close coupled equations
Derivation → blackboard
4.4 Reactive scattering: concepts, methods and examples

4.4.1 Motion on the PES


The topology of the Born-Oppenheimer PES determines the dynamics of a
chemical reaction. Even in the absence of exact reactive-scattering
calculations, important qualitative insight into chemical dynamics can be gained
from inspecting the properties of the PES.
Consider the simplest polyatomic case: the reaction between an atom A and a
diatom BC: A + BC → AB + C .

products
Reaction profile for a linear • The path of minimum energy from the
approach of the reactants
reactants to the products of the PES is
termed reaction path or reaction
saddle point =
transition state coordinate.
• The energy barrier (saddle point on
the surface) separating reactant and
reactants product “valleys” is termed transition
state.
If the total energy in the reactants (the sum of
collisional energy Ec, vibrational energy Ev,
rotational energy Er, and electronic energy Ee
if applicable) is higher than the barrier height,
the reaction can proceed in principle. The
available energy Eavl after the collision is
distributed among the products.

For an A + BC reaction, the barrier height in


general changes for different approach
angles. If more energy is stored in the
reactants, the barrier can also be crossed for
approach angels differing from the optimal
value. Thus, the cone of acceptance of the
reaction can be increased.

Potential energy profile along the reaction coordinate


for H + H2 for different values of the approach angle γ.
P. Siegbahn et al., J. Chem. Phys. 68 (1978), 2457
D.G. Truhlar et al., J. Chem. Phys. 68 (1978), 2466
4.4.2 Effect of vibrational and kinetic energy: Polanyi rules

For asymmetric reactions, the Forward reaction Backward reaction


transition state is usually located closer HF + H → H2 + F H2 + F → HF + H
late barrier early barrier
to either the reactant or the products
(early or late barrier). From an H2 + F HF + H

inspection of the favourable reaction


trajectories, it can be seen that:
For an early barrier, translational HF + H
excitation (high kinetic energies) of H2 + F
the reactants promotes the reaction
and leads to vibrationally excited
products. Vibrational excitation
hinders the reaction. H2 + F HF + H

For a late barrier, vibrational


excitation promotes the reaction
and leads to products with a high
kinetic energy. Translational HF + H H2 + F
excitation of the reactants hinders
the reaction.
4.4.3 Angular momentum constraints

Angular momentum (AM) conservation for the collision dictates:


rotational AM collisional AM
total AM

J� = �jBC + L
� = �jAB
� ��
+L
before collision after collision

If the reactants are internally cold (e.g., from supersonic cooling in a molecular
beam), then the initial rotational AM can be neglected:
J� ≈ L
� = �jAB
� ��
+L
In addition, for reactions involving the transfer of a light atom L from a heavy
atom H’ to another heavy atom H (H + LH’ → HL + H’), we get
J� ≈ L
� ≈L��
because the large rotational energy spacing of HL suppresses rotational excitation
of the product so that orbital AM is conserved. This is called the kinematic effect.
Conversely, for a heavy-atom transfer H + LH’ → HH’ + L we obtain
J� ≈ L
� ≈ �jAB

because the product orbital AM L� = µ� v � b� is usually small owing to the small


reduced mass μ’ of the products. Thus reactant orbital AM is converted into
product rotational AM.
4.4.4 Reaction mechanisms from angular scattering
The angular distribution of scattering products
reflecting the differential scattering cross section can
be measured in crossed molecular beam experiments.
The angular distribution of the scattering products is
measured with a moveable detector in the laboratory
frame. The distribution of scattering angles θ and
product velocities uAB in the centre-of-mass (COM) Schematic of a crossed
frame can be inferred from a Newton diagram (velocity molecular beam experiment
diagram).

Notation:

vA, vBC ... velocity vectors of reactants in lab frame


vrel ... relative velocity vector of the reactants
Θ ... scattering angle in the lab frame
vCM ... velocity vector of the COM
vAB ... velocity vector of product AB in lab frame
uAB ... velocity vector of product AB in COM frame
θ ... scattering angle of products in COM frame

Reconstruction of the COM


Newton diagram for the
angular distribution from a
reaction A + BC → AB + C
CMB measurement
The reconstructed COM product flux distribution ICM(θ,u) can be decomposed
into two different components:

ICM (θ, u) = T (θ) × P (Et� )

product angular distribution product translational energy distribution


(kinetic energy release)

The COM product flux distributions are


usually represented in a polar plot. The
contour lines indicate the product flux
scattered into a certain angle θ with a given
velocity u (or kinetic energy Et’). Example: Product flux distribution
for the HCl product in the reaction
H2 + Cl → HCl + H.

The reaction mechanism manifests itself directly in the angular distribution of


the reaction products. Two important types of mechanisms can be distinguished:
• Direct mechanisms entail a direct scattering event
• Indirect (or complex-forming) mechanisms entail the formation of an
intermediary reaction complex
4.4.4.1 Direct reactions

Two important limiting cases:

• Stripping reactions: dominated by


long-range interactions between the
reaction partners. Occur at large
impact parameters, lead to forward
scattering, i.e., the product angular
distribution peaks at θ=0°. (For A +
BC, “forward” is defined with
respect to the direction of the
incoming atom A.)

• Rebound reactions: dominated by


short-range interactions. Occur at
small impact parameters, lead to
backward scattering, i.e., the
product angular distribution peaks
at θ=180°.
Example I: Cl + H2 → HCl + H
Classical reaction showing rebound
dynamics with a highly constrained
linear transition state. The small cone
of acceptance leads to small impact
parameters and backward scattering.

P. Casavecchia, Rep. Prog. Phys. 63 (2000), 355


M. Alagia et al., Science 273 (1996), 1519
sis offers a way to circumvent this difficulty by
rating the c.m. angular distribution, since 9.7 for Br2 and 12. The halogen polarizabili ties were
estimated from the HX and H2 values 35 via
Qr = 211'
1" (dQr)
-
o dw abs
.
sm()d() (2)

rtue ofExample II: K +


the cylindrical Br2 → KBr
symmetry about+ Br
the initial o K + Br,
ve velocity vector. The absolute normalization of G Rb + Br,

Reaction
ifferential reactiveinitiated
scatteringby long-range
cross section, • Cs + Br,
electron transfer from K to Br2 at a
(dQr/ d<,;) abs =:n( dQr/ dw )rel, (3)
crossing between potential curves
e determined by comparison
corresponding to thewith the elastic
neutral andscat-
,
g. Theionic
resultsforms
obtained fromreactants
of the three different pro-
(harpoon o K + I,
es are given in Table III.
mechanism). The temporary ion pair • Cs + I,

is strongly accelerated
Method A towards one
ce theanother
relative by the Coulomb
intensity scales forinteraction
the reactive
ultimately
lastic scattering areleading to the
practically formation of
the same,3!
00 L-L--'--'---'---L-L....J---"----,-L.L--'---'---'---"----'-I..--J
the products. (dQe/Large
dw) abs impact O· 30· 60· 900 1200 1500 1800
parameters, forward. scattering. (4) CM SCATTERING ANGLE e
( dQe/ dw) reI
FIG. 16. Comparison of approximate C.m. angular distributions
elastic scattering pattern at narrow angles is as- of reactive scattering. The curves (--) are calculated from the
Legendre polynomial expansions given in Table II.
d to be negligibly perturbed by reaction. The ------
ute intensity thus can be calibrated by use of the 33It is expected that dQ,/dw and Q, as predicted from the S--K
approximation shouldcurve correct within 20% (including allow-
be crossing
-angle scattering formula for a VCr) = -e/r 6 van ance for the uncertainty in the polarizabilities). This is indicated
Waals interaction,32 by extensive data on relative cross sections (see Ref. 24) and
recent absolute measurements for several reference systems; see
E. W. Rothe and R. H. Neynaber, J. Chern. Phys. 42, 3306
(1965); ibid. 43, 4177 (1965); and H. G. Bennewitz and H. D.
his is ensured by the data reduction procedure used (rela- Dohmann, Z. Phygik 182, 524 (1965). Small angle scattering
tensity J. H. Birley et al., J. Chem. Phys. 47 (1967), 993
defined by ratio of signal to parent-beam attenua-
D. Hershbach, Angew. Chemie Int. Ed. 26 (2987), 1221
measurements of Ref. 27 (h) give <:l=870XlO--60 erg·cm 6 for
Pt data normalized to W data), provided that: (1) the de- K + Br2, in good agreement with the S--K result of Table III.
4.4.4.2 Indirect reactions
Indirect reactions proceed via the formation of a long-lived reaction complex
(corresponding to a reaction intermediate, i.e., a minimum on the PES along
the reaction path) which lives longer than several rotational periods. During
this time, the collision partners lose part or all of the memory of their original
orientation (see also section 4.4.3):
If L≈L’, i.e., the collisional angular L≈J=L’
momentum and thus the plane of
collision is conserved, the
products show a distinct forward-
backward scattered distribution:
dσ dσ 1
= ∝
dΩ 2π sin θdθ sin θ
L≈J=j’
If L≈j’, i.e., the products are
rotationally excited, memory of
the original orientation is
completely lost and the angular
distribution is isotropic (i.e.,
constant).
Example I: OH + CO → CO2 + H
The reaction of CO + OH (a major channel for the production of CO2 in
combustion processes) proceeds via the formation of an intermediate HOCO
product. The angular distribution shows prominent forward-backward scattering
peaks indicating the indirect mechanism with a propensity for the conservation
of collisional AM.

P. Casavecchia, Rep. Prog. Phys. 63 (2000), 355


M. Alagia et al., J. Chem. Phys. 98 (1993), 8341
Example II: angular product distribution and reaction paths: O(1D) + H2 → OH + H

This reaction can either direct mechanism


proceed through an indirect
insertion mechanism of the O
atom into the H-H bond
forming an intermediary backward scattering
water molecule which breaks
apart or by a direct
abstraction mechanism via an
excited electronic state.
Depending on the collision
energy, both pathways can be
open and can be
distinguished by their
different angular product rotationally excited products:
isotropic angular distribution
distributions.

indirect mechanism
4.4.4.3 Dynamics at curve crossings: adiabatic and diabatic states
Charge-transfer mediated reactions such as
the harpooning reaction in K + Br2 are
classical examples of reactions dominated by
the crossing of two potential energy surfaces.
In fact, many chemical processes are
dominated by such non-adiabatic dynamics
when the system crosses from one PES to
another. Such processes involve a breakdown of Surface crossing in a
the Born-Oppenheimer approximation. charge-transfer mediated reaction

The crossing from one PES to another necessitates coupling terms in the
molecular Hamiltonian which are usually neglected in the BO approximation, e.g.,
• the adiabatic correction terms Ĉn (see section 2.3) which couple states of
the same symmetry and the same multiplicity
• spin-orbit interaction which couples states with different multiplicities (see
problem sheet 3)
Although usually small, such couplings become important when two electronic
states come close in energy, i.e., at crossing points.
Mathematical description:

• Let Φ1(0) and Φ2(0) be electronic states in the BO approximation (so-called


diabatic states), i.e., solutions of a BO-Hamiltonian Ĥ0 (see section 2.2). If these
sates are coupled by an additional weak coupling operator V, the total
Hamiltonian is given by
Ĥ = Ĥ0 + V̂
• The coupled states can be expressed as a superposition of the uncoupled states:
(0) (0)
φ = c1 φ1 + c2 φ2
with mixing coefficients c1 and c2.
• By inserting into the nuclear Schrödinger equation ĤΦ=EΦ, multiplying from
the left bey either Φ1(0) or Φ2(0) and integrating over the nuclear coordinates
(see section 2.3) we get a set of secular equations for c1 and c2:
c1 (H11 − U) + c2 H12 = 0
c1 H12 + c2 (H22 − U) = 0

where Hij = �φ(0)


i | Ĥ|φ
(0)
j � = �φ
(0)
i | Ĥ 0 + V̂ |φ
(0)
j � are the matrix elements of Ĥ in the
diabatic basis.
• Note that
(i) Hii≡Ui(R) (i=1,2), the BO energies of Φ1(0) and Φ2(0)
(ii) H12≡V12(R)
because Φ1(0) and Φ2(0) are orthonormal eigenstates of Ĥ0 and V mixes Φ1(0)
and Φ2(0).

• Note also that both, the BO energies Ui and couplings V12 generally
depend on the reaction coordinate R.

• The secular equations thus become:


c1 (U1 − U) + c2 V12 = 0
c1 V12 + c2 (U2 − U) = 0

• The solutions (energies of the coupled electronic states) are:



1 1
U± (R) = 2 (U1 (R) − U2 (R)) ± 2 (U1 (R) − U2 (R))2 + 4V12 (R)2

with the associated eigenfunctions Φ+ and Φ- (the so-called adiabatic states).


U(R)
The coupling repels the states around adiabatic states

the crossing point and leads to an


avoided crossing.
diabatic states
At the crossing point, the separation
between the adiabatic states is given by
ΔU=2V12.
The adiabatic states and the associated R
PES are the eigenfunctions of the full
Hamiltonian Ĥ and can be obtained Diabatic and adiabatic states at a crossing point

from ab-initio calculations.


In a diatomic molecule, states of the same
symmetry can never cross because of non-
adiabatic couplings. All such crossings are always
avoided (non-crossing rule). This restriction is
relaxed in polyatomics.
Q1
The crossing of two states is referred to as a
conical intersection. The term originates from Q2
diabatic passage
the shape of the two potential energy surfaces
in the crossing region in two dimensions (2D adiabatic passage

cut through the PES along two internal coords Conical intersection between two
Q1 and Q2). electronic states in two dimensions
Conical intersections dominate the dynamics of many chemical processes
involving excited electronic states (see several examples in this chapter).
Moreover, in many cases energy barriers on an adiabatic PES are caused by
avoided crossings.
Landau-Zener theory: When a crossing is traversed in the course of a reaction,
the system can stay on the same adiabatic surface (adiabatic passage) or cross
to the other adiabatic surface (i.e., stay on the same diabatic surface, diabatic
passage).
The probability Pad for diabatic passage (i.e., crossing from one adiabatic surface
to the other) can be calculated using the semiclassical Landau-Zener equation:
� �
2
2πV12
Pdia = exp −
hv ∂(U2 (R)−U
∂R
1 (R))

where v ... velocity along reaction coordinate


U1(R), U2(R) ... BO-PES associated with the diabatic states Φ1(0) and Φ2(0)

The probability for adiabatic passage Pdia is then Pad = 1 − Pdia


Thus, the probability for diabatic passage is high if the coupling V12 is weak and
the velocity and the difference of the potential gradients are large.
Thus, the probability for adiabatic passage is high if the coupling V12 is strong
and the velocity and the difference of the potential gradients are small.
4.4.4.4 Reaction resonances

Reaction resonances are a distinctly quantum


collision
mechanical phenomenon which lead to strong energy Ec
fluctuations in the reaction cross section and

Energy V / ev
Reactants Products
the collision time. They appear when the
collision energy is in resonance with a suitable
bound state of the system thus enhancing the
reaction probability.

Reaction resonances can modulate the


reaction cross section by several orders of
magnitude in a small energy interval. They can

Reaction probability

Reaction time delay


therefore have drastic effects on the
dynamics of a reaction.
There are two important types of reactive
resonance effects:

• Feshbach resonances: the bound state is Collision energy Ec


an excited state of the system (e.g.,
rotationally, vibrationally or electronically
excited)
• Shape (or orbiting) resonances: the
J.N. Milstein et al., New J. Phys. 5 (2003) 52
bound state is located behind a
Example for a dynamic situation leading to a
centrifugal barrier Feshbach resonance

Obviously, the occurrence of resonances


strongly depends on the collision energy,
collisional angular momentum and quantum
state of the reactants. Collision energy Ec

Example for a shape resonance


Example I: F + H2 (j=0) → HF (v’) + H
F + H2 shows a strong forward-scattering peak in the angular distribution of the
HF (v’=3) and HF (v’=2) product around Ec=2.18 kJ mol-1. At this energy,
Feshbach resonances with bound states in the H...HF (v’=3) van-der-Waals
potential well exist which enhance the reaction probability and modify the
product angular distributions. The HF (v’=2) product is then formed by strong
vibrational (anharmonic) coupling between the v’=3 and v’=2 states in H...HF.
Bound van-der-Waals states

Potential well of
H...HF
van-der-Waals
complex

Vibrationally adiabatic
potential curves for
vibrationally excited
states v’ in the HF
product

M. Qiu et al., Science 311 (2006), 1440


X. Wang et al., PNAS 105 (2008), 6227
−0.01 −0.01

−0.02 −0.02
0 5 10 15 20 25 30 35
Example II: Cl + HD (v=1, j=0) → HCl + D / DCl + H
R (au)

Wave functions of the quasibound levels B and E supported by the adiabatic potential shown
The Cl + HD (v=1,j=0) reaction is predicted to have pronounced rotational
as functions of the atom–molecule separation. Amplitudes of the wave functions have been
a factor of 10 for practical plotting reasons.
Feshbach resonances caused by bound states of the van-der-Waals complex
Cl...HD of the reactants.
Long-range interactions in chemical reactions
102
0.7 rotationally adiabatic potential curves
Cl+HD - nonreactive E v=1, j=1
101 DCl+H - reactive
HCl+D - reactive D
100 C
Cross section (10−16 cm2)

0.69 B
10−1 collision
v=1, j=0 energy

Energy (eV)
10−2 A

0.68
10−3

10−4
E
10−5 reactive 0.67
resonances BC D
10−6
10−8 10−7 10−6 10−5 10−4 10−3 10−2 10−1
5 10 15 20
Incident
Collisionkinetic energy
energy (eV)(eV) R (au)
The same as in figure 5 but plotted as a function of the incident kinetic energy to illustrate the
Figure 6.shallow
Adiabaticpotential-energy
potential energy curveswells
of the Cl þ HD system correlating to the HDðv ¼ 1,
ature behaviour of the cross-sections.
causedHDðvby ¼ 1,long-range
j ¼ 1Þ levels as(van-der-Waals)
functions of the atom–molecule separation, R. Quasibound levels re
interactions
for the resonances observed in figure 5 are labelled by B, C, D and E.
Weck and Balakrishnan, Int. Rev. Phys. Chem. 283 (25), 2006
REPORTS

4.4.5 A case study: the SN2 reaction Cl - +details


Until recently, many
CH
namics of bimolecular anion-molecule →I +
of the S 2 dy- - may be obtained from measurements of correlated
3I reactions CH
angle- and 3 Cl N
energy-differential cross sections.
could only be obtained from chemical dynamics Specifically, the probabilities for energy redistri-
simulations. However, with recent experimental bution within the ion-dipole complexes, their
J. Mikosch et al., Science 319(22),
advances (2008),
insight184
into the reaction dynamics dependences on initial quantum states, the branch-

SN2 nucleophilic substitution reactions


X- + R-Y → Y- + R-X show a characteristic transition state

double-well potential-energy profile


along the reaction coordinate.
According to the conventional picture,
the reaction proceeds via a back-side
attack on the R-Y bond leading to an reaction complex in
inversion of the molecular configuration. reaction complex in
exit channel

entrance channel
For the model reaction Cl- + CH3I → I- +
CH3Cl, one would expect that the
dynamics is dominated by the formation Fig. Reaction
1. Calculated profile for
MP2(fc)/ECP/aug-cc-pVDZ Born-Oppenheimer → I- +energy
Cl- + CH3Ipotential −
CH3along
Cl the reaction
coordinate g = R − R for the S 2 reaction Cl + CH I and obtained stationary points. The reported
C−I C−Cl N 3

of a long-lived reaction complex in the energies do not include zero-point energies. Values in brackets are from (28).
exit channel.
If the lifetime of the reaction complex is longer than several rotational
periods, an isotropic product angular distribution is expected.
A B
coordinate g = RC−I − RC−Cl for the SN2 reaction Cl− + CH3I and obtained stationary points. The reported pulsed-field velocity servation
map imaging advances (22), insight into the r
spectrometer,
of energy and momentum (24). of the outermost ring in the image. Thus, the ren

Downloaded from www.sciencem


energies do not include zero-point energies. Values in brackets are from (28). which maps the velocityThe of top I − product
therow anionmaps of the I −
of Fig. 2 shows largest fraction of the available energy is parti-
product ion velocities from the Cl− + CH3I → tioned to internal rovibrational energy of the en
CH3Cl + I − reaction at four different relative col- CH3Cl product. rea
lision energies between Erel = 0.39 eV and Erel = A distinctly different reaction mechanism be- of
1.90 eV, which were chosen because they span the comes dominant at the higher relative collision dis
distinct reaction dynamics observed in this energy energy of 0.76 eV (Fig. 2B): The I − product is the
Isotropic distribution: complex- Forward-scattered products: Forward-backward-scattered products:
range. The only data processing applied to the ion back-scattered into a small cone of scattering ph
mediated, classical mechanism direct substitution mechanism roundabout
impact position on the detectormechanism
is a linear conver- angles. This pattern indicates that direct nucleo- wh
sion from position to ion speed and a transforma- philic displacement dominates. The Cl− reactant trib
tion into the center of mass frame. Consequently, attacks the methyl iodide molecule at the concave do
the velocity vectors of the two reactants, the Cl− center of the CH3 umbrella and thereby drives the agr

A B C anion and the CH3I neutral,D line up horizontally I product away on the opposite side. The direct the
and point in opposite directions, indicated by the mechanism leads to product ion velocities close ve
arrows in Fig. 2. Each velocity image represents a to the kinematic cutoff. In addition, part of the ser
histogram summed over 105 to 106 scattering product flux is found at small product velocities fin
events. The total energy available to the reaction with an almost isotropic angular distribution, in- rep
products is given by the relative translational dicating that for some of the collisions there is a tio
energy, Erel, of the reactants plus the exoergicity, significant probability of forming a long-lived (13
0.55 eV, of the reaction (Fig. 1). I − products reach
Cl- complex. CH3 - I en
the highest velocity when all the available energy At a collision energy of 1.07 eV (Fig. 2C), the CH
is converted to translational energy. The outer- complex-mediated reaction channel is not 0.7
most circle in Fig. 2 represents this kinematic observed any more. The reaction proceeds almost en
cutoff for the velocity distribution. The other con- exclusively by the direct mechanism, with a similar can
centric rings display spheres of the same trans- velocity and a slightly narrower angular distribu- the
lational energy and hence also the same internal tion relative to the 0.76-eV case. At an even val
Angular product
E distributions forF I- at different collision
G energies E ≡E productrel c spacedHat 0.5-eV intervals.
excitation, higher collision energy of 1.90 eV, the domi- in
0.5

In the gas-phase crossed-molecular beam


(40
mi
rel

scattering experiment, three types of product sig


co

angular distribution T(θ) are observed of


for

indicating three different reaction mechanisms: EC


co
(27
Fig. 2. (A to D) Center-of-mass images of the I− reaction product velocity The energy transfer distributions extracted from the images in (A) to (D) in

gie
Isotropic T(θ) at low collision energies E
from the reaction of Cl− with CH3I at four different relative collision energies.
The image intensity is proportional to [(d3s)/(dvx dvy dvz)]: Isotropic scat-
comparison with a phase spacectheory calculation (red curve). The arrows in
(H) indicate the average Q value obtained from the direct chemical dynamics
ag
coe

indicating the classic mechanism via a long-


tering results in a homogeneous ion distribution on the detector. (E to H) simulations. are
at
en

184 lived reactive complex.


11 JANUARY 2008 VOL 319 SCIENCE www.sciencemag.org
bil
1.9


vib
Forward-scattered scattered I- (w.r.t. to res
me

incoming Cl-) indicating a fast, direct Fig. 3. View of a typical trajectory for the indirect roundabout reaction mechanism at 1.9 eV that
Representation of the roundabout the
proceeds via CH3 rotation. ab
nucleophilic displacement of the I-. reaction mechanism

www.sciencemag.org SCIENCE VOL 319 11 JANUAR
Additional forward-backward-scattered I--
products at highest Ec indicate a new indirect
“roundabout” reaction mechanism. Fig. 1. Calculated MP2(fc)/ECP
Fig. 1. Calculated MP2(fc)/ECP/aug-cc
4.5 Photodissociation dynamics and laser chemistry

Chemical processes of molecules excited by light are of relevance for a range of


environments and applications, e.g.,:
• Photochemistry: study and control of chemical reactions by radiation
• Atmospheric chemistry
• Interstellar chemistry
• Radiation damage to biological molecules

In general, the following properties are of relevance for the photodissociation


dynamics of molecules:
• The dissociation energy of the molecule D0
• The symmetries of the involved electronic states
• The absorption cross sections for photoexcitation
• Timescales for the dissociation event
• Product yields if more than one dissociation channel is open
• Angular distributions of the photofragments
4.5.1 Dynamics of electronically excited states

A molecule which is electronically excited by (laser) radiation can undergo a


range of dynamical processes:

a) Laser-induced fluorescence
b) Excitation to the repulsive wall of a bound state, leading to direct dissociation
c) Excitation of a repulsive state, leading to direct dissociation
d) Excitation to a bound state and dissociation by coupling to a repulsive state
e) Excitation to a bound state and dissociation by tunneling through a barrier
f) Excitation to a bound state and dissociation by internal conversion to the
dissociation continuum of the ground state
Processes d)-f) are referred to as predissociation.
4.5.2 Models for photodissociation

Schematic representation of important limiting cases of photodissociation dynamics

Potential
curves of
electronic
states

Quantum states of region of strong coupling between


photofragments different electronic states

a) Adiabatic model: the molecule follows a single potential energy curve during
fragmentation. Applicable if the recoupling region is traversed very slowly.
b) Sudden (diabatic) model: the dissociative molecular states are directly mapped
onto the fragment states. Fragment state distributions are determined by the
overlap of the molecular with the fragment wavefunctions and symmetry/angular
momentum constraints. Applicable if the recoupling region is traversed very fast.
c) Statistical model: all accessible fragment states are equally populated.
Applicable in the limit of very strong coupling between electronic states.
4.5.2 Models for photodissociation

Schematic representation of important limiting cases of photodissociation dynamics

Potential
curves of
electronic
states

Quantum states of region of strong coupling between


photofragments different electronic states

d) Cartoon corresponding to a realistic situation with mixed dynamics


e) Transition state model: dynamics is dominated by a single transition state. In the
statistical limit in which the energy is distributed over all accessible molecular
states, this situation can be described by transition state theory (see, e.g.,
lecture PC IV). Often a good representation of the photodissociation dynamics
in polyatomics.

How can we determine experimentally which situation applies ?


4.5.3 Experimental methods

4.5.3.1 Photofragment translational spectroscopy


PTS is an important method to unravel the energetics and product state
distribution in a photodissociation event AB + hν → A + B.
The total kinetic energy release Et’ (or KER)
of the photofragments A and B is given by
(neglecting internal and kinetic energy of AB
which are usually small in comparison):
Et � = hν − D0 − Eint,A − Eint,B
where Eint,A and Eint,B are the internal =Et’

energies of the fragments A and B. Their


kinetic energies are given by momentum
conservation:
mB mA
Et,A = Et � Et,B = Et �
mAB mAB
Thus, the total kinetic energy release can be calculated by measuring the
velocity of only one of the fragments. In general, the lighter fragment carries
away most of the kinetic energy.
Example: photodissociation of ozone O3 in the Hartley bands: O3 + hν → O2 + O
• Process relevant for shielding the earth’s surface from cosmic UV radiation
• Energy released causes stratospheric temperature inversion
• Complicated process with different competing reaction channels
• Experiment: study photodissociation at 248 nm using an excimer laser
Thelen et al., J. Chem. Phys. 103 (2001), 7946
Absorption spectrum Potential energy curves

electronically
direct excited fragments
dissociation

predisso-
ciation

fragments
in electronic
ground state

Wavelength / nm
Thelen et al., J. Chem. Phys. 103 (2001), 7946
O2 photofragment translational spectra from O3 photodissociation at 248 nm

• The resolved peaks at low Et’ in the spectrum correspond to O2 (1Δ)


photofragments in well-defined vibrational states v produced by dissociation on
the 1 1B2 surface.
• The broad peak at high Et’ corresponds to unresolved, highly excited vibrational
states in the O2 (3Σ) photofragment by predissociation on the 2 1A1 surface.
direct dissociation:
slow fragments
low vibrational excitation

direct dissociation

predissociation:
fast fragments
high vibrational excitation
fragments were detected state specifically by resonance-
enhanced two-colour ionization via selected spin-rotational
levels of the A 2S+ (v 0 = 0) intermediate state. The resulting

4.5.3.2 Velocity-mapped ion imaging (VMI)

VMI has become a standard method for


measuring both, the KER and the
photofragment angular distributions at the
same time.

After photodissociation, one of the


fragment species is ionized by REMPI. The
expanding Newton sphere (the 3D velocity
distribution) of the fragments is then Experimental setup for VMI
accelerated by electric fields and crushed
onto an ion detector. Fig. 1 Schematic diagram of the experimental setup used for the
velocity-map ion imaging studies.

By a specially designed electrostatic lens This journal is #


c the Owner Societies 2007

system, all molecules with the same


velocity vector are mapped onto the same
spot on the detector.

Eppink and Parker, Rev. Sci. Instrum., 68 (1997), 3477 Newton spheres for the photofragments A and B
~d of
the fragment recoil vector and the electric field vector E pertaining t
the dissociation laser). Similarly the angular distribution is is typical f

The central slice of the Newton sphere


can be reconstructed mathematically
from the raw image, e.g., by an inverse
Abel transformation, or experimentally
by only switching on the detector when
the central slice arrives (slice imaging).
Raw VMI image for Reconstructed central
Fig.photodissociation
2 (a) Symmetrized of raw
NO2and (b)of
slice inverse
the NewtonAbel-transforme
sphere
"1
around
the photolysis of380
NOnm
2 at Eexc/hc = 1056.0 cm
S.J. Matthews (correspond
et al.,
PCCP 9 (2007), 5656
3
P1 fragment channels, respectively. (c) Radial distribution e
The radius of the rings in a VMI image is proportional to the fragment velocity
and therefore contains the same information
5658 | as a photofragment
Phys. Chem. Chem. Phys.,translational
2007, 9, 5656–5663
spectrum.
Fig. 2 (a) Symmetrized raw and (b) inverse Abel-transformed image of the NO 2P1/2 (v00
For initially randomly
the oriented
photolysis ofmolecules,
NO2 at Eexc/hcthe photofragment
= 1056.0 angular
cm"1 (corresponding to Edistribution "1
avl/hc = 855.4 cm
in the laboratory frame
3 is given
P1 fragment by (derivation
channels, respectively. see, e.g.,distribution
(c) Radial Zare, Angular
extractedMomentum):
from the image in
1 � � 1
� 2

T (Θ) = 1 + βP2 (cos Θ) with P2 (cos Θ) = 2 3 cos Θ − 1 ... 2. order Legendre polynomial
4π 5658 | Phys. Chem. Chem. Phys., 2007, 9, 5656–5663

β ... anisotropy parameter (-1≤β≤+2) ε


v
Θ
Θ ... angle between the velocity vector v of
the photofragments and the polarization R
vector ε of the photodissociation laser
� �2
The absorption probability P ∝ �µ ε� will show a maximum for molecules with the
� ·�
transition dipole moment μ oriented parallel to ε. In a diatomic molecule, v is
always parallel to the bond vector R . Thus if ...
J. Chem. Phys., Vol. 114, No. 6, 8 February 2001
• µ� � R� (parallel transition), then T(Θ) will be maximal for �v � �ε (β=+2).
• µ� ⊥ R� (perpendicular transition), then T(Θ) will be maximal for �v ⊥ �ε (β=-1)
β≈-1
β≈2
perpendicular transition
parallel transition

adapted from
E. Wrede et al., J.Chem.Phys. 114 (2001), 2629

If β=0, then T(Θ) is isotropic. In this case the dissociation is slower than several
rotational periods and the information about the original molecular orientation is
lost.

Thus, the value of β contains information about the symmetry of the excited
state (which determines whether the transition is parallel or perpendicular, see
section 2.2) as well as about the timescales of the dissociation process.
Example: Imaging of the photodissociation of IBr: IBr + hν → I + Br
E. Wrede et al., J.Chem.Phys. 114 (2001), 2629

• IBr: Hund’s case a: notation of states: 2S+1|Λ|(|Ω|)


• Parallel transition: Δ Ω=0, perpendicular transition: Δ Ω=±1
• Photodissociation at 440 nm shows two velocity Potential energy curves
J. Chem. Phys., Vol. 114, No. 6, 8 February 2001

components corresponding to the formation of


I+Br and I+Br*
• I+Br: β≈-1: indicates perpendicular transition 440 nm

dissociation via the A, and C states

• I+Br*: β≈2: indicates parallel transition


dissociation via the B state
J. Chem. Phys., Vol. 114, No. 6, 8 February 2001 High resolution ion imaging study of IBr photolysis
Photodissociation images at 440 nm
r
B
I+

r*
B
I+

Raw image
Reconstructed Iodine atom product speed distribution
of iodine
central slice
products
4.6 Real-time studies of reactions: femtochemistry

Bond-breaking processes happen on the timescale of molecular vibrations


(femtoseconds, 10-15 s)
real-time studies require the generation of ultrafast laser pulses
Broadband fs laser excitation usually
leads to the excitation of several
vibrational states at the same time.

The vibrational wavefunctions interfere resulting in the formation of


a localised vibrational wavepacket:
wavepacket vibrational wavefunctions
localised
The wavepacket oscillates back and wavepacket |Ψ|2
forth on the excited-state potential after fs excitation
energy surface with a frequency
corresponding to the vibration that
has been excited

fs pump-probe experiments:
a vibrational wavepaket is generated by
a first fs laser pulse (the pump), the time
evolution of the wavepacket is studied
with a second fs pulse after a variable
delay (the probe)
Example I: real-time observation of molecular vibrations
• Step 1: create a vibrational wavepacket consisting of the v=11-15 states in the
first excited electronic state of Na2 using a 50 fs laser pulse

• Step 2: study the motion of the wavepacket by a probe pulse triggered after a
variable time delay
Na2+ signal intensity

(T. Baumert et al., J. Phys. Chem. 95 (1991), 8103)


Example II: transition state dynamics in NaI
Zewail and co-workers, Annu. Rev. Phys. Chem. 41 (1990), 15

• Consider two lowest electronic states of


NaI with potential energy curves V0(R) and wavepacket
V1(R) motion

• Both states exhibit an avoided crossing at


R=Rc at which they strongly interact.

• A vibrational wavepacket is created in the


excited state by fs laser excitation

• The wavepacket oscillates in the excited-


state potential well. Every time it avoided
approaches the avoided crossing, part of crossing
the population crosses to the ground-state
adiabatic potential curve on which the
molecule dissociates.
• Experiment: probe wavepacket motion with a second fs laser pulse
- at the inner turning point of the excited-state potential (trace b)
- at large internuclear distances on the ground-state surface (trace a)

wavepacket
motion

a b

avoided
crossing

Вам также может понравиться