Sie sind auf Seite 1von 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 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 ?

b)

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 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

b)

a)

Chapter 4: Contents

4.1 Reaction rates and cross sections 4.2 Classical scattering theory 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

b)

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: bimolecular: AB A+BC

The rate v for a bimolecular reaction is given by


dCA v = = k(T )CA CB dt
thermal rate coefficient number 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 factor activation energy

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 (*) = CB I dx
reaction cross section

ln(I1 /I0 ) = CB Integrate: 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: dCA dI d(CA vA ) dCA = = vA = dx dx dx dt Using I=vACA , the right-hand side of Eq. (*) becomes: CB I = vA CA CB

Thus, Eq. (*) becomes: Comparison with (**) yields:

dCA = vA CA CB dt k = vA

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

where

p(E) = E

(kB T )3/2 2

exp

E kB T

... Maxwell-Boltzmann kinetic-energy distribution

Inserting p(E) and using k(E)=v=(2E/)1/2 where is the reduced mass we get: 1 E 3/2 exp k(E) = E (kB T ) (E) 2E/ dE 2 kB T 0 Changing to the dimensionless energy variable (E/kBT) leads to: 1/2 8kB T E dE E k(E) = (E) exp kB T kB T kB T 0
average thermal velocityvrel

This expression can be formulated as where


k(T ) = vrel E dE E = (E) exp ... averaged cross section kB T kB T kB T 0

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 )CA = vrel CB B

The number of collisions between A and B molecules per unit volume is thus:
V ZAB = vrel CA CB

If A=B, we get:

V 2 ZAA = (1/2)vrel CA

(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 0 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: 0 = d 2 = (rA + rB )2 The thermal rate constant is: 1/2 8kB T E dE E k(T ) = 0 exp kB T kB T kB T 0 Evaluation of the integral conveniently yields 1: 1/2 8kB T k(T ) = 0

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: b0: head-on collision b>>0: glancing collision The reaction cross section can generally be formulated as bmax = P (b)2b 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 are usually dependent on the collision energy. If P(b)=1, we recover the hard-sphere collision model: bmax 2 = 2b db = bmax
0

3. Reactions with a threshold (activation) energy E0: Reaction only occurs if E>E0: 1/2 8kB T E dE E k(T ) = 0 exp kB T kB T kB T E0

Integration yields:
k(T ) = 0

8kB T

1/2

E0 1+ kB T

exp

E0 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 E0 8kB T A(T ) = 0 1+ 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 coordinate frame defined by
r r c = mAA + mB B R mA + mB r R = A B r

c.o.m.
B

rA Rc ... coordinate vector of the centre of mass (c.o.m.) ... relative coordinate vector rB x

It can easily be shown that the kinetic energy Ekin of the system is given by:
1 1 1 2 1 2 2 2 Ekin = mA vA + mB vB = MV + v 2 2 2 2 r where the velocities are given by vi = |i |,V = |Rc | , v = |R| mA mB M = mA + mB is the total mass and = 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 conserved, two coordinates suffice to describe the relative motion of the collision partners. We choose R ... relative position vector ... orientation angle of R with respect to the original velocity vector v
impact parameter
je ct or y

initial velocity vector

deflection angle

vector of closest approach relative position vector orientation angle

co lli

si

on

tra

Conserved physical quantities


v Angular momentum: L = R

= R sin v

|L| = L = v b

initial velocity vector

deflection angle

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

vector of closest approach

1 2 1 = R + R2 2 + V (R) with the angular velocity = d/dt = 2 2 1 2 1 L2 = R + + V (R) 2 2 2 R 1 2 1 L2 E = R + + V (R) 2 2 2 R


VL(R) ... centrifugally corrected (effective) potential

L = R2

co

lli

L before collision = L after collision: |L| = L = | R| where v is the initial velocity vector v

si

on

tra

je

ct

or y

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 The effective potential for the collision contains both, the interaction potential V(R) and the centrifugal energy:
1 L2 VL (R) = + V (R) 2 2 R

centrifugal barrier
Centrifugally corrected potentials VL(R) for L3 > L2 > L1 > L0=0

The deflection function (b) The angle of deflection depends on the impact parameter b. Examples: 1. Hard-sphere collisions (a billiard game):

For b > d: = 0 For b < d: = 20 where b/d = sin 0 = = 2 arccos(b/d)

2. General potentials with repulsive and attractive parts:


repulsive, short range part: V(R)>0

b* = b/Re

g
V(R) R*

attractive, long range part: V(R)<0

Small b: collision dominated by repulsive forces Large b: collision dominated by attractive forces

backward scattering forward scattering

Rainbow angle r: maximum negative deflection angle at impact parameter brRe where the potential is most attractive Glory angle g: deflection angle at impact parameter bgR* where attractive and repulsive forces cancel

Experimentally, it is not possible to distinguish between positive and negative deflection angles because of the cylindrical symmetry of the collision process. One can only measure the absolute value of the deflection angle =||.

Experimental observables in molecular-collision experiments The intensity of scattered molecules I(), i.e., the flux of molecules scattered into the solid angle , defines the differential cross section d/d:
d

d scattered ux of molecules per unit solid angle I() = = d incident ux of molecules per unit area

The integral cross section is obtained by integration.


= d d = 2 d
0

d sin d d

Where the cylindrical symmetry of the problem allowed us to express


d = 2 sin d

in the second step. The scattering rate constant is then given by (see section 4.1.2):
k = v

Calculating the differential cross section from the deflection function (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) d = 2b db Again, because the scattering problem is cylindrically symmetric, the solid angle element d can be formulated as
d = 2 sin d

Hence we obtain for the differential cross section: d 2b db b = I() = = d 2 sin d sin (d/db) If more than one value of b contribute to the same scattering angle , we have to sum over all contributions and arrive at the following dependence of the differential cross section on the deflection function (b): d b = I() = d sin (d/db) Singularities in the differential cross section (d/d=):

Glory (=0) singularity: sin = 0 Rainbow singularity: (d/db) = 0 (maximum of the function (b) )

Illustration:
d b = I() = d sin (d/db)

deflection function (b)

differential cross section I()

glory singularity

rainbow singularity

Calculating the deflection function (b) from the potential V(R) It can be shown (see, e.g., R.D. Levine, Molecular Reaction Dynamics):
(b) =

i.e., (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) (b) E

b R2

1 1
b2 R2

V (R) E

1/2 dR

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 .
Reaction profile for a linear approach of the reactants saddle point = transition state

products

The path of minimum energy from the reactants to the products of the PES is termed reaction path or reaction coordinate. The energy barrier (saddle point on the surface) separating reactant and product valleys is termed transition state.

reactants

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 .
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 transition state is usually located closer to either the reactant or the products (early or late barrier). From an inspection of the favourable reaction trajectories, it can be seen that: For an early barrier, translational excitation (high kinetic energies) of the reactants promotes the reaction and leads to vibrationally excited products. Vibrational excitation hinders the reaction. For a late barrier, vibrational excitation promotes the reaction and leads to products with a high kinetic energy. Translational excitation of the reactants hinders the reaction.
Forward reaction HF + H H2 + F late barrier H2 + F Backward reaction H2 + F HF + H early barrier HF + H

HF + H H2 + F

H2 + F

HF + H

HF + H

H2 + F

4.4.3 Angular momentum constraints


Angular momentum (AM) conservation for the collision dictates:
rotational AM total AM collisional AM

j j J = BC + L = AB + 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 J L = AB + 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 J L AB because the product orbital AM L = v b is usually small owing to the small reduced mass 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 and product velocities uAB in the centre-of-mass (COM) frame can be inferred from a Newton diagram (velocity diagram).
Notation: vA, vBC ... velocity vectors of reactants in lab frame vrel ... relative velocity vector of the reactants ... 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 ... scattering angle of products in COM frame Newton diagram for the reaction A + BC AB + C Reconstruction of the COM angular distribution from a CMB measurement

Schematic of a crossed molecular beam experiment

The reconstructed COM product flux distribution ICM(,u) can be decomposed into two different components:
ICM (, u) = T () 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 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 =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 =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

(dQr) sm()d() . (2) o dw abs Example II: K symmetry about the rtue of the cylindrical + Br2 KBr + Br initial ve velocity vector. The absolute normalization of Reaction initiated by long-range ifferential reactive scattering cross section, electron transfer from K to Br2 at a (3) (dQr/ d<,;) abs =:n( dQr/ dw )rel, crossing between potential curves e determined by comparison with the elastic scatcorresponding to the neutral and g. Theionic forms of the reactants (harpoon results obtained from three different proes are given in Table III. mechanism). The temporary ion pair Method A is strongly accelerated towards one ce theanother by the Coulomb interaction relative intensity scales for the reactive ultimately leading to the formation of lastic scattering are practically the same,3! the products. Large impact (dQe/ dw) abs . parameters, forward scattering. (4) ( dQe/ dw) reI Qr = 211'

" 1

9.7 for Br2 and 12. The halogen polarizabili ties were estimated from the HX and H2 values 35 via

o K + Br, G Rb + Br, Cs + Br,

,
o

K + I, Cs + I,

00 L-L--'--'---'---L-L....J---"----,-L.L--'---'---'---"----'-I..--J O 30 60 900 1200 1500 1800 CM SCATTERING ANGLE e

elastic scattering pattern at narrow angles is asd to be negligibly perturbed by reaction. The ute intensity thus can be calibrated by use of the -angle scattering formula for a VCr) = -e/r 6 van Waals interaction,32

FIG. 16. Comparison of approximate C.m. angular distributions of reactive scattering. The curves (--) are calculated from the Legendre polynomial expansions given in Table II.
-----33

his is ensured by the data reduction procedure used (relasignal 47 (1967), 993 tensity J. H. Birley et al., J. Chem. Phys.to Ed. 26 (2987), 1221 defined by ratio ofChemie Int. parent-beam attenuaD. Hershbach, Angew. Pt data normalized to W data), provided that: (1) the de-

It is expected that dQ,/dw and Q, as predicted from the S--K curve correct approximation should be crossing within 20% (including allowance for the uncertainty in the polarizabilities). This is indicated 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. Dohmann, Z. Phygik 182, 524 (1965). Small angle scattering measurements of Ref. 27 (h) give <:l=870XlO--60 ergcm 6 for 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 LL, i.e., the collisional angular momentum and thus the plane of collision is conserved, the products show a distinct forwardbackward scattered distribution:
d d 1 = d 2 sin d sin

LJ=L

LJ=j

If Lj, 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 proceed through an indirect insertion mechanism of the O atom into the H-H bond forming an intermediary 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 distributions.
direct mechanism

backward scattering

rotationally excited products: isotropic angular distribution

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 the Born-Oppenheimer approximation.

Surface crossing in a 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 n (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 1(0) and 2(0) be electronic states in the BO approximation (so-called diabatic states), i.e., solutions of a BO-Hamiltonian 0 (see section 2.2). If these sates are coupled by an additional weak coupling operator V, the total Hamiltonian is given by H = H0 + V The coupled states can be expressed as a superposition of the uncoupled states:
= c1 1 + c2 2
(0) (0)

with mixing coefficients c1 and c2.

By inserting into the nuclear Schrdinger equation =E, multiplying from the left bey either 1(0) or 2(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 = (0) |H|(0) = (0) |H0 + V |(0) are the matrix elements of in the i j i j diabatic basis.

Note that (i) HiiUi(R) (i=1,2), the BO energies of 1(0) and 2(0) (ii) H12V12(R) because 1(0) and 2(0) are orthonormal eigenstates of 0 and V mixes 1(0) and 2(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 V12 + c2 (U2 U) = 0 c1 (U1 U) + c2 V12 = 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 + and - (the so-called adiabatic states).

The coupling repels the states around the crossing point and leads to an avoided crossing.

U(R)

adiabatic states

At the crossing point, the separation between the adiabatic states is given by U=2V12. The adiabatic states and the associated PES are the eigenfunctions of the full Hamiltonian and can be obtained from ab-initio calculations.
R

diabatic states

Diabatic and adiabatic states at a crossing point

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

Q1 Q2
diabatic passage adiabatic passage

Conical intersection between two 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 2V12 Pdia = exp hv (U2 (R)U1 (R)) R
where v ... velocity along reaction coordinate U1(R), U2(R) ... BO-PES associated with the diabatic states 1(0) and 2(0)

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.

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

4.4.4.4 Reaction resonances


Reaction resonances are a distinctly quantum mechanical phenomenon which lead to strong fluctuations in the reaction cross section and 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 therefore have drastic effects on the dynamics of a reaction.

collision energy Ec Energy V / ev Reactants Products

Reaction probability

Reaction time delay

There are two important types of reactive resonance effects:

Feshbach resonances: the bound state is an excited state of the system (e.g., rotationally, vibrationally or electronically excited) Shape (or orbiting) resonances: the bound state is located behind a centrifugal barrier

Collision energy Ec

J.N. Milstein et al., New J. Phys. 5 (2003) 52

Example for a dynamic situation leading to a 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

H...HF

Potential well of 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

Example II: Cl + HD (v=1, j=0) HCl + D / DCl + H


Wave functions of the quasibound levels B and E supported by the adiabatic potential shown as functions of the atommolecule separation. Amplitudes of the wave functions have been a factor of 10 for practical plotting reasons.

10

20 15 R (au)

25

30

0.02 35

The Cl + HD (v=1,j=0) reaction is predicted to have pronounced rotational Feshbach resonances caused by bound states of the van-der-Waals complex Cl...HD of the reactants.
Long-range interactions in chemical reactions
102 101 100 Cl+HD - nonreactive DCl+H - reactive HCl+D - reactive

0.7

rotationally adiabatic potential curves


E D C

v=1, j=1

Cross section (1016 cm2)

101

0.69

B v=1, j=0 A

102 103 104 105 106

Energy (eV)

collision energy

0.68

reactive resonances
107 104 103 106 105 Incident kinetic energy (eV) Collision energy (eV)

E BC D
102 101

0.67

108

10

15 R (au)

20

The same as in figure 5 but plotted as a function of the incident kinetic energy to illustrate the Figure 6. Adiabatic potential energy curveswellsCl HD system correlating to the HDv 1, shallow potential-energy of the ature behaviour of the cross-sections. HDv long-range (van-der-Waals) interactions caused by1, j 1 levels as functions of the atommolecule separation, R. Quasibound levels re 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
N

Until recently, many 2 4.4.5 A case study: the SN2 reaction Cl- +details of 3I reactions - + CH3Cl measurements of correlated CH the Sdy- may be obtained from I angle- and energy-differential cross sections. namics of bimolecular anion-molecule could only be obtained from chemical dynamics simulations. However, with recent experimental J. Mikosch et al., Science 319(22), insight into the reaction dynamics advances (2008), 184

Specifically, the probabilities for energy redistribution within the ion-dipole complexes, their dependences on initial quantum states, the branch-

SN2 nucleophilic substitution reactions X- + R-Y Y- + R-X show a characteristic 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 inversion of the molecular configuration.

transition state

For the model reaction Cl- + CH3I I- + CH3Cl, one would expect that the Reaction profile for Cl- + CH3 I- CH3Cl the reaction MP2(fc)/ECP/aug-cc-pVDZ along dynamics is dominated by the formation Fig. 1. Calculated R for the S 2 reactionBorn-OppenheimerIpotential+energypoints. The reported Cl + CH I and obtained stationary coordinate g = R of a long-lived reaction complex in the energies do not include zero-point energies. Values in brackets are from (28). exit channel.
CI CCl N 3

reaction complex in entrance channel

reaction complex in exit channel

If the lifetime of the reaction complex is longer than several rotational periods, an isotropic product angular distribution is expected.
A B

Downloaded from www.sciencem

coordinate g = RCI RCCl for the SN2 reaction Cl + CH3I and obtained stationary points. The reported energies do not include zero-point energies. Values in brackets are from (28).

pulsed-field velocity servation of energy and momentum (24). map imaging spectrometer, which maps the velocity of top row product shows maps of the I The the I of Fig. 2 anion

Isotropic distribution: complexmediated, classical mechanism

Forward-scattered products: direct substitution mechanism

E F G Angular product distributions for I- at different collision energies E E

of the outermost ring in the image. Thus, the largest fraction of the available energy is partiproduct ion velocities from the Cl + CH3I tioned to internal rovibrational energy of the CH3Cl + I reaction at four different relative col- CH3Cl product. lision energies between Erel = 0.39 eV and Erel = A distinctly different reaction mechanism be1.90 eV, which were chosen because they span the comes dominant at the higher relative collision distinct reaction dynamics observed in this energy Forward-backward-scatteredion energy of 0.76 into (Fig. 2B):cone Iof product is products: eV a small The scattering range. The only data processing applied to the back-scattered roundabout mechanism impact position on the detector is a linear conver- angles. This pattern indicates that direct nucleosion from position to ion speed and a transforma- philic displacement dominates. The Cl reactant tion into the center of mass frame. Consequently, attacks the methyl iodide molecule at the concave the velocity vectors of the two reactants, the Cl center of the CH3 umbrella and thereby drives the anion and the CH3I neutral, line up horizontally I product away on the opposite side. The direct D and point in opposite directions, indicated by the mechanism leads to product ion velocities close arrows in Fig. 2. Each velocity image represents a to the kinematic cutoff. In addition, part of the histogram summed over 105 to 106 scattering product flux is found at small product velocities events. The total energy available to the reaction with an almost isotropic angular distribution, inproducts is given by the relative translational dicating that for some of the collisions there is a energy, Erel, of the reactants plus the exoergicity, significant probability of forming a long-lived 0.55 eV, of the reaction (Fig. 1). I products reach complex. ClCH3 - I At a collision energy of 1.07 eV (Fig. 2C), the the highest velocity when all the available energy is converted to translational energy. The outer- complex-mediated reaction channel is not most circle in Fig. 2 represents this kinematic observed any more. The reaction proceeds almost cutoff for the velocity distribution. The other con- exclusively by the direct mechanism, with a similar centric rings display spheres of the same trans- velocity and a slightly narrower angular distribulational energy and hence also the same internal tion relative to the 0.76-eV case. At an even H higher collision energy of 1.90 eV, the domiproductrel excitation, spaced at 0.5-eV intervals. c

advances (22), insight into the r

ren

In the gas-phase crossed-molecular beam scattering experiment, three types of product angular distribution T() are observed indicating three different reaction mechanisms:
Fig. 2. (A to D) Center-of-mass images of the I reaction product velocity from the reaction of Cl with CH3I at four different relative collision energies. The image intensity is proportional to [(d3s)/(dvx dvy dvz)]: Isotropic scattering results in a homogeneous ion distribution on the detector. (E to H)

en rea of dis the ph wh trib do agr the ve ser fin rep tio (13 en CH 0.7 en can the val in 0.5 (40 mi rel sig co

184

Isotropic T() at low collision energies E indicating the classic mechanism via a longlived reactive complex. 319 SCIENCE www.sciencemag.org 11 JANUARY 2008 VOL Forward-scattered scattered I- (w.r.t. to incoming Cl-) indicating a fast, direct nucleophilic displacement of the I-. Additional forward-backward-scattered I-products at highest Ec indicate a new indirect roundabout reaction mechanism.

The energy transfer distributions extracted from the images in (A) to (D) in comparison with a phase spacec theory calculation (red curve). The arrows in (H) indicate the average Q value obtained from the direct chemical dynamics simulations.

of for EC co (27 gie ag coe are at en bil 1.9 vib res me

Fig. 3. View of a typical trajectory for the indirect roundabout reaction mechanism at 1.9 eV that proceeds via CH3 rotation.

Representation of the roundabout reaction mechanism


www.sciencemag.org SCIENCE

the ab

VOL 319

11 JANUAR

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 photofragments

region of strong coupling between 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 photofragments

region of strong coupling between 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 + h 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 = h D0 Eint,A Eint,B where Eint,A and Eint,B are the internal 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
=Et

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 + h O2 + O

Process relevant for shielding the earths 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 excited fragments

direct dissociation

predissociation

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 (1) 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 (3) 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 specically by resonanceenhanced 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 accelerated by electric fields and crushed onto an ion detector. By a specially designed electrostatic lens system, all molecules with the same velocity vector are mapped onto the same spot on the detector.

Experimental setup for VMI


Fig. 1 Schematic diagram of the experimental setup used for the velocity-map ion imaging studies.

This journal is

 c

the Owner Societies 2007

Eppink and Parker, Rev. Sci. Instrum., 68 (1997), 3477

Newton spheres for the photofragments A and B

~ the fragment recoil vector and the electric eld vector E d of the dissociation laser). Similarly the angular distribution is

pertaining t 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).

Reconstructed central Raw VMI image for Fig.photodissociation of raw2and (b)of the Newton sphere 2 (a) Symmetrized NO slice inverse Abel-transforme S.J. Matthews et al., around NO2 at (correspond the photolysis of380 nm Eexc/hc = 1056.0 9cm15656 PCCP (2007),
3

The radius of the rings in a VMI image is proportional to the fragment velocity and therefore contains the same information| as a photofragment translational 5658 Phys. Chem. Chem. Phys., 2007, 9, 56565663 spectrum.

P1 fragment channels, respectively. (c) Radial distribution e

Fig. 2 (a) Symmetrized raw and (b) inverse Abel-transformed image of the NO 2P1/2 (v00 For initially randomly orientedof NO at E /hc = 1056.0 cm1 (corresponding to Edistribution 1 the photolysis molecules, the photofragment angular avl/hc = 855.4 cm 2 exc 3 in the laboratory frame is given by (derivation see, e.g., Zare, Angular Momentum): in P1 fragment channels, respectively. (c) Radial distribution extracted from the image

1 2 1 P2 (cos ) = 2 3 cos 1 ... 2. order Legendre polynomial T () = 1 + P2 (cos ) with 5658 | Phys. Chem. Chem. Phys., 2007, 9, 56565663 4
... anisotropy parameter (-1+2) ... angle between the velocity vector v of the photofragments and the polarization vector of the photodissociation laser R v

2 The absorption probability P will show a maximum for molecules with the transition dipole moment oriented parallel to . 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 v R (parallel transition), then T() will be maximal for (=+2). v R (perpendicular transition), then T() will be maximal for (=-1)
-1 perpendicular transition 2 parallel transition

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

If =0, then T() 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 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 + h I + Br


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

IBr: Hunds case a: notation of states: 2S+1||(||) Parallel transition: =0, perpendicular transition: =1 Photodissociation at 440 nm shows two velocity components corresponding to the formation of I+Br and I+Br* I+Br: -1: indicates perpendicular transition dissociation via the A, and C states I+Br*: 2: indicates parallel transition dissociation via the B state Photodissociation images at 440 nm
B r
J. Chem. Phys., Vol. 114, No. 6, 8 February 2001

Potential energy curves

440 nm

J. Chem. Phys., Vol. 114, No. 6, 8 February 2001

High resolution ion imaging study of IBr photolysis

I+

Raw image of iodine products

Reconstructed Iodine atom product speed distribution central slice

I+

r*

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

The wavepacket oscillates back and forth on the excited-state potential 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)

localised wavepacket ||2 after fs excitation

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 V1(R) 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 excitedstate potential well. Every time it approaches the avoided crossing, part of the population crosses to the ground-state adiabatic potential curve on which the molecule dissociates.

wavepacket motion

avoided crossing

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

avoided crossing

Das könnte Ihnen auch gefallen