You are on page 1of 36

Chemical kinetics

M F Suárez
*
−E
Z AB =π d 2 v N A N B 2
Z AB =π d v N A N B e kT

*
−E
2 kT
K th T =π d ve  K exp T 
A potential energy surface for a direct unimolecular reaction without a saddle
point. The surface corresponds to a reaction like H2O → H + OH for dissociation
along a fixed bond angle, where only two internuclear coordinates are required
in order to specify the configuration.
A potential energy surface for a direct bimolecular reaction. The surface corresponds to a
reaction like D+H − H → D − H+H at a fixed approach angle, say in a collinear configuration
specified by the D–H and H–H distances. These distances are measured along the two
perpendicular axes. (Note that in this figure all energies above a fixed cut-off value Emax
have been replaced by Emax.)
Chemical Kinetics
Video

Schematic illustration of a classical trajectory (dotted line)


superimposed on the potential energy surface for the reaction
A + BC → AB + C. Note the vibrational energy in BC, and that
this particular trajectory leads to chemical reaction.
Contour plot of the H3 potential energy surface in Cartesian coordinates. The H-H distance of the
reactant H2 molecule is fixed at RBC = 1.044 Å and the potential is plotted as a function of the (x,y)
position of the third H atom. If the third H atom drops into the well at linear geometry it can form a new
bond with the nearest atom of the former molecule while the bond to the remaining atom is broken.
Note that the potential is cylindrically symmetric around the x axis.
Mapping out a reaction. Surface mesh
plot of the potential energy surface for
the collinear H1 + H2H3 → H1H2 + H3
reaction as a function of the two bond
distances, R, superimposed on a
conventional contour plot of the
potential energy surface. The minimum-
energy reaction path is indicated by the
solid orange line. The zero-point energy
in the reactant and product
arrangements, and at the barrier, is
indicated by straight black lines. The
three corresponding green slices
indicate the range of coordinate space
made classically accessible by zero-
point energy.
TS
−Δ U m (0)
qi , m νi
hν0 T
K (T )=e RT

i ( )
o
A
θv =
kB
qv≈
θv

TS
−Δ U (0)
TS (q q ' ) ̊
A
A+B AB → Product [ ABTS ]=[ A][ B] v , RC total −RC AB e k T
q A qB
B

TS
−Δ U (0)
dP Tk q ' ̊
A
=ν ' [ ABTS ]=ν ' [ A][ B] B total − RC , AB e k T B

dt h ν0 q A qB

TS
̊ −ΔkU T (0)
Tk B q ' AB A
K (T )= e B

h qA qB
43 A.
Wassermann, Monatsch. Chem., 83, 543 (1952).
44 A.
U. Blackham and N. L. Eatough, J. Am. Chem. Soc., 84, 2922
(1962).
Comparison of Arrhenius activation
energies (Ea ≡ Earrhenius) calculated
via QM (red solid lines), ICVT (dashed
black lines), and ICVT/LAG (solid
black lines) for the reactions of 4.1 H
and 0.11H with H2 as a function of
temperature (T ).

LAG includes tunneling


Science 28 January 2005: Vol.
307 no. 5709 pp. 558-563

Schematic of the potential energy


landscape of ground and excited
states in complex molecular systems.
The equilibrium ground-state structure
defines the initial wave packet prepared
by femtosecond pulse excitation to the
excited-state surface. Because of its
nonequilibrium nature, the excited state
structure evolves into radiative and
radiationless channels. The radiationless
transitions can result from bifurcation
into reactive chemical processes and
nonreactive physical pathways (internal
conversion/ intersystem crossing).
The primary effect of quantum mechanical tunneling on chemical
kinetics is that it can be seen deviations from classical kinetic
behavior

2a
Particle Mass l (Å)
(a.m.u)
e- 1/1750 26,9
h
H 1 0,63 =
mv
D 2 0,45
C 12 0,18
Br 80 0,07

l: De Broglie wavelength
kinetic energy = 20kJ/mol m: mass
v: velocity
Consequences of Tunneling
on Reaction Kinetics

R.P. Bell developed a quantum tunneling correction factor, Q, and explored its
effect on an Arrhenius treatment of reaction kinetics
−Δ U eα
K (T )=QAexp (
RT ) Q=
β−α
−α −β
[β e −α e ]
1/2
ΔU 2 (2m ΔU )
α= β=2a π
RT h

This equation relates measurable reaction parameters to the probability of


tunneling, allowing us to experimentally determine if tunneling is taking place.

Four key experimental observations that imply tunneling is taking place

1. Large Kinetic Isotope Effect


2. Temperature Independence
3. Anomalous A values
4. Anomalous Ea values
Large KIE
Classical kinetics: kH / kD arises from
difference in ZPE's
(EaD – EaH) max ≤ 1.354 kcal / mol

Temp ( ̊C) kH / kD

-30 17
-100 53
-150 260
In organic reactions, this proton tunneling effect has been observed in
such reactions as the deprotonation and iodination of nitropropane with
hindered pyridine base with a reported KIE of 25 at 25 °C:

Edward Sheldon Lewis and Lance Funderburk (1967). "Rates and isotope effects in the proton transfers from 2-
nitropropane to pyridine bases". J. Am. Chem. Soc. 89 (10): 2322–2327. doi:10.1021/ja00986a013.
Reaction of HO with CO:
Tunneling Is Indeed Important

J. Phys. Chem. Lett. 2012, 3, 1549−1553

Most important reaction paths on the lowest-lying doublet


electronic potential energy surface in the HO + CO reaction,
constructed using the HEAT protocol.
Calculated thermal rate
constants as a function of
temperature at the zero-
and high-pressure limits
for the HO + CO reaction.
Experimental data are
included for the purpose
of comparison.
Modelo que representa la transferencia de carga entre dos
moléculas cuyas coordenadas internas son q1 y q2.
Cl oxidation video

Following ILET by Polarization Anisotropy: Femtosecond polarization anisotropy methods are used to
follow the hopping of the electron from ligand-to-ligand (i.e. interligand electron transfer or ILET) that
takes place in the metal-to-ligand charge transfer (MLCT) excited states of two Os(II) dyes. Adiabatic
ILET in the Os(bpy)32+ complex occurs with an 8 ps hopping time in acetonitrile.
J. Am. Chem. SOC.,
Vol. 106, No. 10, 1984

Intramolecular electron-transfer
rate constants as a function
of free energy change in MTHF
solution at 296 K. Electrons
transferred from biphenyl ions to
the eight different acceptor
groups, A (shown labeling the
points), in eight bifunctional
molecules having the general
structure shown in the center of
the figure.
Curva de energía libre diabática para una reacción de
transferencia de carga asimétrica.
1 1
V R ( qc )=V P ( qc ) f (q c −q R,0 )2 =∆ G 0 + f (q c −q P ,0 )2
2 2

∆ G0 1 1
qc = + (q P ,0 +q R,0 ) ∆ G ‡= f (q c −q R ,0 )2
f (q P ,0 −q R ,0 ) 2 2

2
1 ( ∆ G 0+ λ )
λ= f (q P ,0−q R ,0 )2 ∆ G ‡=
2 4λ

{ }{ }
2
1 1 1 1 1
( )
0
−( ∆G + λ ) λ 0= ( Δe )
2
+ − −
k ET = A Arrhenius exp 2 a1 2 a2 R ϵ ∞ ε0
4 λ kBT
0 2 2
( ∆G + λ ) =(−eη+ λ ) =W
dW
W ≈ W ( η=0 ) +
dη ( ) η=0
2
( η−0 ) =λ −2 λeη

−λ 2 −2 λe η
(
k ET = AArrhenius exp
4 λ k BT ) (
= A Arrhenius exp
−λ−2 eη
4k BT )

k ET =k exp
0
2 kBT (
=k ox
)
Electron flow through proteins and protein assemblies in
the respiratory and photosynthetic machinery commonly
occurs between redox active cofactors that are
separated by large molecular distances, often on the
order of 10–25 Å. Although these cofactors are weakly
coupled electronically, the reactions are remarkably rapid
and specific. Understanding the underlying physics and
chemistry of these distant electron transfer processes
has been an overarching goal of theorists and
experimentalists for many years.

−√8m r Δ E
k ∝exp( )

Angew. Chem. Int. Ed.2013,52,1–5
Collision between two spheres:

Since there are no internal degrees of


freedom that may exchange energy with the
translational degrees of freedom, the
collision is elastic, i.e., the total kinetic
energy of the two particles before and after
the collision is conserved.
Video about collision theory
SN2

Laura Pedraza-González, Johan F. Galindo, Ronald González, and Andrés Reyes* Departamento de Quıımica, Universidad Nacional de
Colombia, Av. Cra 30 # 45-03, Bogotá, ColombiaJ. Phys. Chem. A, 2016, 120 (42), pp 8360–8368DOI: 10.1021/acs.jpca.6b06517Publication
Date (Web): October 9, 2016
Diels-Alder Reactions
J.Org.Chem.2014, 79, 8968−8976
Phys. Chem. Chem. Phys., 2016,18, 31895-31903

You might also like