Peter C. Jordan Chemical Kinetics
Peter C. Jordan Chemical Kinetics
and Transport
Chemical Kinetics
and Transport
Peter
C. Jordan
Brandeis University
Waltham, MasSllchusetts
78-20999
Preface
This book began as a program of self-education. While teaching undergraduate physical chemistry, I became progressively more dissatisfied with
my approach to chemical kinetics. The solution to my problem was to
write a detailed set of lecture notes which covered more material, in greater
depth, than could be presented in undergraduate physical chemistry. These
notes are the foundation upon which this book is built.
My background led me to view chemical kinetics as closely related to
transport phenomena. While the relationship of these topics is well known,
it is often ignored, except for brief discussions of irreversible thermodynamics. In fact, the physics underlying such apparently dissimilar processes
as reaction and energy transfer is not so very different. The intermolecular
potential is to transport what the potential-energy surface is to reactivity.
Instead of beginning the sections devoted to chemical kinetics with a
discussion of various theories, I have chosen to treat phenomenology and
mechanism first. In this way the essential unity of kinetic arguments, whether
applied to gas-phase or solution-phase reaction, can be emphasized. Theories
of rate constants and of chemical dynamics are treated last, so that their
strengths and weaknesses may be more clearly highlighted.
The book is designed for students in their senior year or first year of
graduate school. A year of undergraduate physical chemistry is essential
preparation. While further exposure to chemical thermodynamics, statistical
thermodynamics, or molecular spectroscopy is an asset, it is not necessary.
As a result of my perspective, the choice of topics and organization of
the material differs from that of other kinetics texts. The introductory
chapter, on equilibrium kinetic theory, is mainly review. The next two
chapters, treating transport in both gas and solution, introduce many
v
vi
Preface
Contents
Notation . . . . . . . . . . . . . . . . . . . . . . . . . . . ..
1.
2.
1
2
6
9
13
14
15
3.
xi
17
18
20
24
26
29
30
35
44
45
47
Introduction . . . . . . . . . .
Ionic Conduction in Solution-a Review.
vii
49
50
viii
Contents
3.3. Nernst-Einstein Relation
3.4. Electrolyte Diffusion
3.5. Stokes' Law and the Microscopic Interpretation of AD
3.6. The Mobility of H+ and OH- in Water
3.7. The Concentration Dependence of A
3.8. The Wien Effects
Problems . . . . .
General References
4.
69
5.
53
56
57
61
62
65
67
71
72
73
77
78
80
85
87
89
93
98
101
103
11 0
113
114
120
124
127
131
135
139
142
145
147
153
156
157
163
ix
Contents
6.
Photochemistry
6.1.
Introduction
7.
Introduction . . . . . . . .
8.
165
166
167
169
178
184
185
188
192
195
199
201
202
205
209
214
220
226
229
232
Single-Collision Chemistry
8.1.
Introduction . . . . .
233
234
237
240
244
248
252
254
257
259
261
263
267
Contents
x
9.
Physical Constants
Index . . . . . .
325
326
328
333
335
339
342
345
348
351
353
355
357
Notation
Those symbols which are only used once in the text do not appear in this
listing. Familiar mathematical functions are also not included, nor are
quantities that are simply intermediates in a derivation. Capital letters are
listed before lowercase letters. Greek letters are listed at the end.
A
ABC
Ci
Cv
D
D
Di
FA(C)
E
Ea
E
F
FA(u, v, w)
FA(C)
L1Go+
G{J,(u)
Notation
xii
iJHo+
Enthalpy of activation
h
J(v)
J(J.., x)
JA(x)
J
Ji
Jq
K(T), K(T, p), K(T, p, I)
K+
Km
M
MA
No
NA
Nt
N( E,,)
Px
P( E)
Q
QA
Q+
Q(E, i,j), Q(E), Q(E)
Si
iJ So t
T
Ti
VA
molecule
Intensity of a beam of molecules of species A
Normalized fluorescence intensity at frequency v
Position-dependent light intensity at wavelength J..
Position-dependent intensity of a beam of molecules of species A
Rotational quantum number
Diffusive flux of the ith component
Heat flux
Equilibrium constants
Equilibrium constant for activation
Michaelis constant
Total mass
Molecular weight of species A
Avogadro's number
Number of particles of species .A
Hypothetical number of activated complexes
Density of states for the ath degree(s) of freedom
of an activated molecule with energy E~ in the
specified degree(s) of freedom
Momentum flux in x direction
Probability density for states of energy E
Enthalpy release per mole of reaction
Molar partition function for species A
Reduced molar partition function for the activated complex
Reaction cross sections as functions of molecular
states i, j, collision energy E, or molar collision
energy E
Differential cross section for molecules with
collision energy E scattering into solid angle 1>
Gas constant, total reaction rate
Rate of forward (reverse) reaction
The ith singlet state of an electronic manifold
Entropy of activation
Absolute temperature
The ith triplet state of an electronic manifold
Potential energy of a molecule of species A
e,
Notation
xiii
V
Vcoll
L1 vot
X, Y
Yi
[Yd
ZA
ai
at
b
C
Co
Ci
Ci
CM
tv
d (dAB)
e
erf(x)
fA (c)
f
g~
h
ji
k
ken), kn
kd
kdiff
(k diss )
koo (k o)
kef:)
me
mi
xiv
Notation
ni
qA
q+
qrx
r
r rx
,i
'M
u
u(r)
v
w
x, y
x, y, z
Zi
ZA(C)
rex)
L1
'l'
complex
Pressure
Partial pressure of the ith component
Position-dependent scattering probability for
molecules in a beam
Reaction probability as functions of molecular
states i, j, collision energy E, and impact parameter b
Molecular partition function for species A
Reduced molecular partition function for the
activated complex
Partition function for the ath independent
degree(s) of freedom
Distance in spherical polar coordinates
Rate of the ath step in a reaction sequence
Position of the ith particle
Position of the center of mass
Time
x component of velocity, mobility in an electric
field
Intermolecular potential function
y component of velocity, volume per mole,
reaction velocity in an enzyme-catalyzed reaction, vibrational quantum number
Drift velocity in an external field
z component of velocity
Position components for the equivalent particle
in the dyIiamical model of Chapter 10
Position components in Cartesian coordinates
Charge of the ith ionic species
Speed-dependent collision frequency per molecule of species A
Gamma function
Reaction exo- or endoergicity for dynamical
model of Chapter 10
Equivalent conductance of the ith electrolyte
(at infinite dilution)
Time- and position-dependent wave function
xv
Notation
ai(A)
fJ
Y
Yi
y+
Y(A)
+
.j(v)
1]
()
()i
"
"i
AA
Ai (AiO)
fli
fli
flij
v
vA
Vi
Notation
XVI
~(E)
ei
ei(E)
eiu)
a
r(A)
cp
CP;.
CPI (cpQ)
X
1p
Kinetic Theory
of Gases-Equilibrium
1.1. Introduction
The kinetic theory of gases represents the first truly successful effort
to construct a model which provides a mechanical basis for understanding
the properties of bulk matter. By combining the molecular hypothesis
with Newtonian mechanics and using a statistical approach, the theory
provides an explanation for the thermodynamic similarities between dilute
gases, makes numerous verifiable predictions, and suggests a number of;
unifying principles. In its simplest form the theory is based upon the following premises:
(1) a gas is composed of an enormous number of molecules;
(2) in the absence of external forces particles move in straight lines(l);
(3) particles collide infrequently;
(4) collisions, whether with the walls or with other molecules, are
elastic (momentum and kinetic energy are conserved).
The interparticle forces actually cause trajectories to bend as molecules approach one
another. However, quite accurate predictions follow if billiard ball dynamics are assumed.
Chap. 1
Sec. 1.2
1=
-00
-00
(1.1)
-00
(1.2)
f +OO eiu) du
-00
(1.3)
f(c)
(l.4)
(5)
The concept is analogous to the familiar idea of spatial density. The number of molecules in a volume element limited by x and x + dx, y and y + dy, and z and
z + dz is dN = e(x, y, z) dx dy dz; e(x, y, z) is the number density, i.e., the number
per unit volume.
J. C. Maxwell, Phil. Mag. 19, 19 (1860).
Chap. 1
ex(u)evCv)ez(w)
(1.5)
( a In F(u, v,
au
din ex(u)
du
w) )
11,W
u dlnf(e)
e
de
(1.6)
din exCu)
du
d Inf(e)
de
(1.7)
dlnf(e)
de
-2Ae
which is
Inf(e)
-Ae 2
+ In B
or in exponential form
fee)
B exp( -Ae2 )
(1.8)
The rationale for the negative sign in the exponential is now clear; in
this way infinite velocities are forbidden. Factoring (1.8), the separate
distributions in u, v, and w can be obtained; ex(u) is
(1.9)
with similar expressions for ey(v) and ezCw). Evaluation of A and B requires
further information. The normalization (1.1) or (1.3) provides one condition; in the next section the other is found.
Before determining A and B let us derive (1.8) in a way that does not
presume the factorization property (1.5). Consider collisions in a dilute
(6)
A more sophisticated derivation, not based upon (1.5), was formulated a few years later
[J. C. Maxwell, Trans. Roy. Soc. 157, 49 (1867)]. It is presented later in this section.
Sec. 1.2
gas of identical molecules. As interparticle distances are large only a relatively few molecules are interacting at any instant; of these interactions
the overwhelming majority involve pairs of particles. The trajectories of a
colliding pair are determined by the intermolecular forces between them.
As long as the forces are short range the details of the force law are unimportant. (7) The physical significance of such collisions in dilute gases
is that they provide a mechanism for momentum transfer between molecules.
In a typical collision molecules approach with initial velocities CI and C2 ,
interact, and then separate with final velocities c l ' and c2'. The details of
the interaction need not concern us; we only need to know the initial and
final velocities. Limiting consideration to structureless identical particles,
energy conservation requires that(S)
(1.10)
For the process considered the rate must be proportional to the fraction
of molecules of each velocity, i.e., to F(c l )F(c 2 ). The rate of the reverse
process, in which molecules initially with velocities c/ and c2 ' emerge with
velocities CI and C2 , is analogously proportional to F( cl')F( c2 '). In a system
at equilibrium the rate of any process and its inverse are equal. (9) Since,
as we shall now demonstrate, the proportionality constants determining
the two rates are equal, detailed balance requires that
(1.11)
which, with (1.10), has the unique solution (1.8).
To establish that the proportionality constants are the same, consider
the collision processes as viewed by an observer moving with the centerof-mass velocity (c I + c 2 )/2 = (c l ' + c2 ')/2. Combining this with (1.10)
we find that the relative speed is the same before and after collision,
I C I - C 2 I = I c/ - c 2 ' I = C 12 Collision accomplishes nothing more than
(7)
(8)
(9)
.r;;,
Chap. 1
(0)
(b)
(e)
(d)
Sec. 1.3
Determination of Pressure
J-x
Fig. 1.2. Collision cylinders for molecules with different velocities VI and V.,
but the same x component of velocity,
striking a target area A in a time .dt.
between u and u + du that strike an area A in a time LIt. From (1.2) the
total number of such molecules is Nex(u) duo Only a small fraction of these
hit the wall during a limited time. From Fig. 1.2 we see that the molecules
within the x distance u LIt hit the wall during LIt. The values of v and w
are irrelevant. (10) The fraction that hits the designated area is the ratio of
volume, Au LIt, of the collision cylinder to the volume, V, of the container,
so that the total number of molecules providing a momentum transfer
2mu to the area A in the time LI t is
Au LIt
- V - Nex(u) du
(1.13)
Integrating over all u greater than zero (only molecules moving toward
the wall may collide) yields the total momentum transfer at x = L; dividing
by A and LIt we obtain the pressure
p
Since eiu)
(10)
= ex(-u)
VN
2m
foo u2ex(u) du
0
Molecules within this distance may collide with other molecules before striking the wall.
Although kinetic energy and momentum are conserved (collisions are elastic) each
molecule will generally change its x component of velocity. Simultaneously (i.e., during
.dt) molecules with different x components of velocity will collide and one of the collision products may then have an x component u. The rate of these processes must
be equal at equilibrium. Otherwise the distribution i!x(u) would be time dependent,
which, as we have seen, it is not. Thus, on balance, for each molecule with component u initially within u.dt of the wall, a molecule (not necessarily the same one) with
component u will strike the wall during .dt.
Chap. 1
f::
u2ex(u) du
Nmu 2
(1.14)
= jNE = jE
(1.15)
where E is the average kinetic energy per molecule and E the total kinetic
energy. Dilute gases are ideal; thus pV = NkT, which, with (1.15), yields
two fundamental results relating mean kinetic energy and mean molecular
speed to temperature,
tmC2 =
ikT,
E= iNkT
(1.16)
The first of these defines the root mean square (rms) speed
Crms
= (C 2)1I2 = (3kTlm)1/2
(1.17)
Bl/3
p VI N = kT =
f~ exp( -
BlI3
Au 2 ) du
f~ mu2 exp( -
(1.18)
Au 2) du
2(AB+l)1I2
0OO
US exp( -Au2) du
rex
+ 1) =
J'' O X(S-1)/2 r
0
dx
S + 1 )
== r (--2-
(1.19)
r function are
r(1) = 1,
xr(x),
(1.20)
+ 1) =
r(n + t) =
r(n
(11)
nt,
(2n)!
0,1,2, ...
n
(1.21)
Sec. 1.4
B1I3 = (m/2nkT)1I2
Now, using (1.8) and (1.9) the velocity distribution function is completely
specified:
(lx(U) = (m/2nkT)1I2 exp( -mu 2/2kT)
(1.22)
F(c) = F(u, v, w) = (m/2nkT)3!2 exp( -mc 2/2kT)
2.5
200
400
600
800
u (ms')
Fig. 1.3. One-dimensional velocity distribution function for N2 at 100 K and 300 K.
Chap. 1
10
(1.23)
F(c)
(1.24)
This distribution is plotted in Fig. 1.4 in terms of the dimensionless parameter c(M/2RT)1/2. Such a plot adjusts the speed scale to account for effects
due to mass and temperature variation.
The maximum in the speed distribution, which occurs at the most
probable speed cmp (M/2RT)1/2 = 1, corresponds to speeds of 422 m S-1
and 244 m S-1 for N2 at 300 K and 100 K, respectively; plotted on the same
scale the speed distribution at 100 K would be more sharply peaked than
that at 300 K, just as was the case for the one-dimensional velocity distribution of Fig. 1.3. The maximum is due to two opposing effects. As the speed
increases, the exponential factor f( c) decreases, reflecting the consideration
that a higher speed means a higher energy; on the other hand the factor
4nc 2 increases, reflecting the fact that higher speeds can be attained in many
2
cmed
3
cJM/2RT
Fig. 1.4. Speed distribution F(c) as a function of dimensionless speed. Various measures
of molecular speed are indicated.
Sec. 1.4
11
more ways than lower ones. (12) The result is the most probable speed,
which is found by calculating dF(c)/dc and setting it equal to zero. This
measure of molecular speed is somewhat different from the rms speed
already calculated, another consequence of having a distribution. A third
common measure of speed is the average (or mean) speed, c, given by the
cF(c) dc; using the formulas (1.19)-(1.21) yields
integral
f:
c=
(SRT/nM)1I2
(1.25)
CI
= c/cmp =
c(M/2RT)1/2, the
(1.26)
2 fX exp( _y 2 ) dy
--m
F or many purposes the fraction with speeds above the cutoff c I is of interest.
Integrating by parts and setting Yu =
the result is
(12)
(13)
f:
exp( _y 2 ) dy]
--m
Yl exp( -YI2) +
n
[1 - erf(YI)]
(1.27)
Molecules with speed c have their velocity vector c on the surface of a sphere of
surface area 4:n:c 2 The larger the value of c the larger this area is.
This function can be looked up in mathematical tables; it is defined so that erf(O) = 0
and erf(oo) = 1. Using these tables is little more trouble than using ordinary tables
of trigonometric functions. See M. Abramovitz and I. A. Stegun, eds., Handbook of
Mathematical Functions (New York: Dover Publications Inc., 1965), p. 310.
Chap. 1
12
which can be used to provide yet another measure of molecular speed, the
median speed; by definition half the molecules move faster and half slower
than the median. To compute the median refer to tables of the error function
and find the value of yz for which (1.27) equals 0.5; the result is YI = 1.0876
or Cmed = 1.538(RT/ M)1I2. The relative values of the four measures of
molecular speed are then
Cmp : Cmed :
C:
Crms =
which are also indicated in the plot of F(c). When the cutoff speed is large,
(1.27) is simplified. For 10% accuracy the integral term can be ignored
whenever Y I ;C 3; if 1% accuracy is required Y I ;C 7.
The dimensionless parameter which characterizes the speed distribution
is the ratio of the molecular kinetic energy to kT. Thus we may determine
a Maxwellian energy distribution by defining E = mc 2/2 from which dE =
mc dc and (1.24) becomes
F(e) de
= F(E) dE
2n(l/nkT)3/2E1I2e-<lkT dE
(1.28)
which measures the fraction of molecules with energy between E and E + dE.
A plot of F(E) would be similar to that of F(c) in Fig. 1.4. Of more
interest is the fraction of molecules with energies greater than a specified
value Eo. This quantity, F*(Eo), is foc
F(E) dE. Integrating by parts the
<0
result, expressed in terms of a dimensionless parameter a o = Eo/kT, is
(1.29)
Sec. 1.5
13
lOOK
Fig. 1.5. Fraction of molecules with energies greater than Eo. The fractions with energy
greater than 6.9 x 10- 21 J at various temperatures are indicated.
A LltG(u) du
(Au LltJV)Nez(u) du
G(u) du
(NJV)uez(u) du
or
(1.30)
For effusion into vacuum there is no return flow and the total flux is found
by integrating over all values of u greater than zero (only molecules with
Chap. 1
14
these velocities are moving toward the hole). Using (1.22) and (1.19)-(1.21)
the total flux is
G=
VN foo0 uex(u) du =
nc/4
(1.31)
with n = N/ V. If the area of a pinhole can be measured, an effusion experiment provides a direct measure of the mean speed; if not, effusion
can be used to compare the mean speeds of different gases, (14)
GI
nIcI
7J; = n2 c2
nl (TI /m l )1/2
n2(T2/m 2 )1/2
(1.32)
Problems
1.1.
1.2.
1.3. A I-liter cube contains N2 at 300 K and 10- 3 atm. Using the uncertainty
principle compute the minimum uncertainty in the specification of x component of velocity u. Accepting' this quantity as the minimum possible
value of .d u, compute the fraction of molecules with x component of velocity
between u and u + .du for (a) u = ii = 0, (b) for u = Urms = (kT/m)1/2.
How many molecules are in the two velocity domains?
1.4.
Show that the speed distribution (1.24) and the energy distribution (1.28)
are both normalized.
1.5.
1.6.
(14)
The effect was put to practical use in the Manhattan project. In order to construct an
atomic bomb 235U had to be separated from its far more abundant isotope 238U. This
was originally accomplished by building a huge gas-effusion system for the separation
of gaseous 2lI5UF. from .38UF.
15
General References
1.7.
Integrate F(c, q;) dc dq; over all q; and show that the two-dimensional speed
distribution is
F(c) dc
c,
1.8.
1.9.
Cmed,
and crma ?
D: 0.00028
23BU: 0.99285
pure 23U. How many passes through an effusion apparatus are required
to enrich naturally occurring uranium to 95% 235U? to 99% 235U? The
effusing gas is UF 6
General References
s.
Kinetic Theory
of Gases-Transport
2.1. Introduction
According to the kinetic model, molecules move very rapidly. At 300 K
the rms speed of gaseous bromine is 125 m sec-I. Yet, if a small amount of
bromine is added to nitrogen at 300 K and I atm the rate at which the
bromine diffuses through the gas is much slower, about 10-3 m sec-I. The
presence of the nitrogen impedes the movement of the bromine, reducing
the effective rate of travel by a factor of 105 from that which occurs in the
absence of nitrogen. The most obvious possible cause for such retardation
is the effect of molecular collision.
As long as the dimensions of the container are large enough so that
molecules undergo collisions in traversing the vessel, a simple, although
imperfect, analogy is found in the new amusement park ride of miniature
hovercraft supported above a rink by air jets. Imagine that all drivers are
blindfolded, the steering wheels of their craft locked so that they move in
straight lines, and that they are required to cross the arena. If there are
few craft the chances are excellent that no collisions will occur and that
they will traverse the floor without incident. As more craft participate
collisions become more likely and the distance traveled between them decreases. Each collision redirects the craft, which goes charging off on its
new path almost surely to collide with another vehicle. The possibility of
making headway is enormously reduced and the time required to cross
the floor increases greatly. On the molecular level the difficulties are much
the same. Thus, the time required for a molecule to traverse a container
17
Chap. 2
18
Direction
of be om of
A molecules
AX
of gas is related to the distance traveled between collisions or, equally well,
to the collision frequency.
-h(x)p(x) dx
(2.1)
(1)
The hard-sphere assumption is a considerable restriction, and most certainly an unrealistic one. See footnote (1) in Chapter l.
Sec. 2.2
19
dlA
dx = - Th(x)
A ]
f x~oo [--dl
-- X
x~O
h(O)
fool_A_ _ X dx =
f oo p(x) [leX)]
o
IA(O)
e-xllx dx = I
(2.4)
If the target gas were a mixture of A and B the analysis would have to
be generalized to account for A-A collisions. Then p(x) and I would no
longer be given by (2.2). Assuming the target gas mixture is dilute enough
so that the collision probabilities are additive the appropriate generalization is
(2.5)
with
d AA = 2rA.
20
Chap. 2
As previously noted, collisions alter the speed of the individual molecules. However,
as the gas is at equilibrium, the total number of A molecules with a given speed is
invariant. It is thus correct to consider the properties of molecules of a particular
speed even though the members of this class are continually changing. See Section
1.2 and footnote (10) in Chapter 1.
Sec. 2.3
Collision Frequency
21
+ drp.
Chap. 2
22
1.0
0.8
2
nnd >-(0.)
0.6
0.4
ex. = c.JM/2RT
Fig. 2.3. (nnd')}.(a), the ratio of the speed-dependent mean free path to its value at
infinite speed, in a one-component system as a function of the scaled speed u.
Sec. 2.3
Collision Frequency
23
The integral in (2.14), which is the mean relative speed CAB, is most
easily evaluated by transforming to center-of-mass and relative coordinates,
(2.15)
To reexpress the product FA( cl)FB( c 2) note that the exponential factor
is dependent upon the total kinetic energy of the two particles, (mAc 12 +
mnc 22)/2. If (2.15) is solved for Cl and C2 and these values substituted in the
expression for the total kinetic energy, it takes the form (MCM 2 + fJc 2)/2,
where M is rnA + mB, and fJ, the reduced mass for an A-B pair, is mAmB/M.
Using these expressions along with (1.22) we find that
FA(c l )FB(C2)
)3/2( 2nkT
fJ )3/2
exp[-(MCM2 + fJc 2)/2kT]
M
( 2nkT
FM(CM)FR(C)
(2.16)
Note that the transformed representation is a product of Maxwellian distributions of the center-of-mass and relative kinetic energy. To complete
the transformation of (2.14) requires expressing dC l dC 2 in terms of the new
coordinates. The result, obtained by the Jacobian method outlined in the
Appendix, is dC l dC 2 = dCM dc so that the integral in (2.14) is the product
FM(CM) dCM
FR(C)C dc
+ mB)/nmAmB]1I2
(2.18)
Note that CAB is always larger than the mean speed of either molecule.
The mean collision frequency for an A molecule in a binary mixture
is thus
(2.19)
since cAA is 2l / 2CA. When combined with (2.13) the mean free path is
approximated by
(2.20)
24
Chap. 2
Sec. 2.4
25
-D dnA
(2.23)
dz
Tz
Fig. 2.4. Velocity profile in a gas undergoing viscous flow. The lower plate
is stationary while the upper one has a
velocity Vx = t5U.
Moving plate
6U
--------~--------+.~=
------+.
v. = z6U/L
------------- v = 0
Stationary plate
26
Chap. 2
differ significantly, so does the mean speed of each one. There is no reason
to expect that the flux of each is the same. In fact, the temperature gradient
leads to a flux of matter, the phenomenon of thermal diffusion. (4) Similarly
a concentration gradient causes heat flow as well as the matter flow demanded
by Fick's law.
Rigorous phenomenological justification of transport equations is
the domain of irreversible thermodynamics. (5) By extending the entropy
concept to systems slightly displaced from equilibrium and requiring that
entropy increases with time, it is possible to generalize flux-force relations
like (2.22)-(2.24) to account for coupled phenomena such as thermal
diffusion. An important consequence is that there is an inherent correlation
between reciprocal processes such as mass flux due to a temperature gradient
and heat flux due to a concentration gradient. The transport coefficient
for the process and its inverse are not independent. Such constraints on the
transport coefficient for a process and its inverse are quite general and are
known as Onsager reciprocal reiations.(6)
Sh(z) - Sh(z
+ LIz)
Since the volume is S LIz the number density of A is NA/S LIz and because
(4)
(5)
(6)
Thermal diffusion has been used as a basis for isotope separation techniques [K. Clusius and O. Dickel, Naturwiss. 26, 546 (1938)].
For an introduction to this subject see I. Prigogine, Thermodynamics of Irreversible
Processes, 3rd ed. (New York: Interscience, 1967).
The theorems demonstrating the interconnections were first proved by L. Onsager,
Phys. Rev. 37, 405 (1931); 38, 2265 (1931).
Sec. 2.5
27
Regions initially
containing
pure B
Region to which
A was originally
confined
J(z
+ LIz) =
J(z)
+ LIz aJ/az
(2.25)
ae )
( a2T)
(at
z =" az
t
(2.26)
aCE/V) )
at
E/ V,
aCE/V) ) (~)
aT
at
~ (~)
V
at
at
z =
" ( a 2T )
az 2 t
cv! V
(2.27)
28
Chap. 2
The three equations (2.25), (2.27), and (2.28) all have the same form.
While general solutions for arbitrary boundary conditions present complicated mathematical problems, (7) certain one-dimensional problems are
simple and instructive. Consider a cylinder of which one end is held at
temperature T* and the material inside is initially at temperature To.
This is a model for heat flow in a long metal rod or in a tube containing
liquid or gas. Making suitable correspondences it also describes diffusion
in a long tube where the solute concentration at one end is kept constant.
Ignoring effects due to the surface of the cylinder, the boundary conditions
are
z = 0, all t
T= {T*,
(2.29)
z> 0, t = 0
To,
The solution to (2.29), which may be verified by direct differentiation, is
T- To _
T* _ To -
2
n 1/ 2
f= exp( -y ) dy -
2_
1 - erf(s)
(2.30)
(8)
For general solutions see J. Crank, The Mathematics of Diffusion (Oxford: Clarendon
Press, 1975).
J. Crank, The Mathematics of Diffusion (Oxford: Clarendon Press, 1975), p. 87.
Sec. 2.6
29
Z2
A point 0.3 m from the end of the rod reaches 100C in approximately
193 sec or a little over 3 min. (9)
The solution to the diffusion equation (2.30) implies that the temperature decreases
monotonically with z at all times. It is interesting that this uniform behavior is not
a property of ultrapure materials. Instead, heat is propagated as a wave, in a fashion
similar to the propagation of sound. This phenomenon, known as second sound, was first
discovered in liquid helium at very low temperatures. It has since been observed in solid
helium and has been reported to occur as well in the isotopically and chemically pure
materials NaF and Bi at very low temperatures. For a discussion of the effect see B.
Bertram and D. J. Sandiford, Scientific American 222(5), 92 (1970).
30
Chap. 2
(11)
For consideration of transport when A > L see M. Knudsen, The Kinetic Theory of
Gases (New York: John Wiley and Sons, 1950).
This is not quite accurate; there are small perturbations to the distribution function
induced by the macroscopic gradients. To treat such effects properly requires a far
more elaborate treatment than is given here. See R. D. Present, Kinetic Theory of Gases
(New York: McGraw-HilI, 1958) for details.
Sec. 2.7
31
z
Direction of
macroscopic
gradients
Direction of
macroscopic
fluxes
--~~~----~-------x
Fig. 2.6. Coordinate geometry for determining the differential molecular flux through a
reference plane at zo.
in the vicinity of a reference plane located at Zo. The further dr is from zo,
the greater the differences. If molecules from dr reach Zo without undergoing
a collision they will on average, contribute to the macroscopic flux through
Zo in a manner determined by the density, momentum, and temperature
in dr. Of course the further dr is from Zo the greater is the chance of a
collision which leads to reequilibration in a new region. We assume that
such reequilibration is established with each collision, i.e., that molecules
are only aware of the last stimulus to which they have been exposed. Thus
to compute the flux of any property at Zo requires determination of the
differential flux of molecules originally in dr that reach Zo without an intermediate collision. The differential flux multiplied by the mean local energy
or momentum yields the differential energy or momentum flux. Then
summing over all regions dr yields the corresponding macroscopic flux.
The total number of particles in dr is n(z) dr, where n(z) is the number
density, which may be position dependent. Of these only the fraction
moving in the direction of r can pass through the specified area A in the
reference plane Zo. Those molecules originally in dr that have not undergone
a collision spread out uniformly over a sphere of area 4nr2. Since the projection of A onto this sphere is A 1cos X I, the fraction which could pass
through A is A cos X 1/4nr2. Collisions reduce the flux through A. In
Section 2.2 we found that the fractional change in flux intensity per unit
distance traveled is A-Ie-riA. Then, since molecules, on the average, travel
a distance eLl t during an interval Ll t, the total number of molecules initially
in dr that pass through A in a time t without undergoing an intermediate
collision is
n(z) dr(A cos X 1/4nr2)(A- I e- rIA )c Llt
1
Chap. 2
32
Since molecules crossing the reference plane from above are moving in the
-z direction the differential molecular flux at Zo is
dJ = -n(z)(cos xI4nr 2A)ce-
r, ). dr
(2.31)
When 0 < X < nl2 (molecules arriving from above zo) the differential
flux is negative; when nl2 < X < n it is positive. (12)
Application of (2.31) is straightforward. To obtain Poiseuille's law
compute the momentum flux for a system in which the mean value of the
x component of velocity is position dependent,
Px
(2.32)
f mii(z) dJ
ii(zo)
+ z -dii
- I + -Z2
dz
2
zo
d 2ii
dz
-2-
0 we find
I + ...
Zo
(2.33)
Since viscous flow occurs at constant temperature and density, the momentum flux, with the geometry of Fig. 2.6, is
mnc
cos
P = - - foo f" f2" ( -X
- e- rl )' )
x
4:n;}..
0
0
0
r2
X
( iio
iio"
+ ... ) r2 sin X dr dX dw
x-
mnAc dii I
- - - 3 - dz
(2.34)
Zo
n
(12)
mnU
3
_2_ (nmkT)1I2
3n
nd 2
(2.35)
It is possible to modify (2.31) to account for the distribution of molecular speeds and
thereby to consider the speed dependence of the mean free path. However, such modification does not incorporate the effect of macroscopic gradients on the distribution
function. Furthermore, it is still restricted to billiard ball dynamics. Thus, even though
it is conceptually more in tune with the previous analysis we do not pursue it here.
Sec. 2.7
33
(2.36)
E(Z) dJ
where E(Z) is the z-dependent mean energy. Since heat flow occurs in a
system at constant pressure, the density is not uniform, n(z) = p/kT(z).
Furthermore the mean speed is also position dependent, C = [8kT(z)/nm]1I2.
Finally the mean energy of an ideal gas is cvT(z), where Cv is the constant
volume heat capacity per molecule. Substituting into the heat flow integral
(2.36), expanding T(z) in a Taylor's series about zo, and integrating, the
result is
(2.37)
which is Fourier's law (2.22). The thermal conductivity may be identified as
nCvAC
6
Cv (nmkT)1/2
-'-----,='-m
nd 2
" = - - - = - - -3n
(2.38)
The simple mean free path approach is not adequate for describing
mutual diffusion in a binary system at constant temperature. Only in
one case is the theory self-consistent-when the molecules are mechanically
indistinguishable (self-diffusion). By integrating (2.31) we obtain the total
flux of one component, h = f dh. Since the only position-dependent
quantity is nA(z), the analysis used in the derivation of Poiseuille's law
yields
(2.39)
precisely Fick's law (2.23) from which we identify
(2.40)
The same arguments may be applied to component B; the analogs to (2.39)
and (2.40) are obtained. From (2.23) we note that there is only a single
diffusion coefficient, i.e., DA = DB. In general AACA
ABCB so that the
*-
(13)
This expression for Ais used for internal consistency. Its derivation, like this treatment
of viscosity, did not consider the variation of mean free path with speed.
Chap. 2
34
mnD
Section 2.6
Exact
0.159
0.375
0.159
0.3125
0.0796cv
m'X
0.3125cv
+ 0.7031k
2
3n
(nmkT)1!2
nmnd 2
= - - -'-----;:::0--
(2.41)
where n = nA + nB.
Our treatment is limited to a very simple model, a dilute gas of hard
spheres. Nonetheless the results are only approximate. The expressions
for'Y) and D, (2.35) and (2.41), differ from the results of an exact theory of
hard-sphere transport by a numerical factor. The expression for :;c, (2.38),
is in error in an additional way. The derivation does not account for possible
differences in the rate of transport of translational energy (center-of-mass
kinetic energy) and of rotational and vibrational energy (internal energy).
Exact theory does; the results, assuming internal energy is also equilibrated
with each collision, are summarized in Table 2.1.
Both treatments lead to similar qualitative conclusions. The viscosity
of a gas is predicted to be independent of the density, an astonishing result
when first discovered by Maxwell. Experiment has confirmed the theory,
which is readily understood qualitatively. As the density increases the
molecular flow increases proportionately. On the other hand, the mean free
path decreases and collisions occur more often. The two effects exactly
(14)
For a discussion of this point see J. E. Mayer and M. G. Mayer, Statistical Mechanics
(New York: John Wiley, 1940), p. 30.
Sec. 2.8
35
balance. Another surprise was the prediction that gas viscosity increases
with temperature, a result opposite to the more familiar situation in liquids,
where viscosity decreases rapidly as temperature increases.
For the same qualitative reasons K, like 'f}, is density independent. The
thermal conductivity is more temperature sensitive than the viscosity
because of the dependence on heat capacity. Its behavior is also quite
different from that observed in liquids. In gases K increases with T, while
in liquids the opposite behavior is normal.
Unlike 'f} or K, D is density dependent since diffusion, as discussed in
the introduction to this chapter, is essentially a direct measure of the mean
free path. The temperature dependence is a consequence of the fact that
faster molecules diffuse more rapidly. From Table 2.1 we see that the ratio
mnDj'f} is predicted to be constant in dilute gases; this is completely different
from Walden's rule in liquids, D'f} f':::i const.
(O.4cv
mK
mnD = 1.2
'f}
(2.42)
25
(2.43)
+ O.9k)'f)
are predicted to be constants, independent of temperature or of the particular gas. These ratios are tabulated for a number of dilute gases at
Table 2.2. Exact Theoretical Expressions for Transport Coefficients of Hard
Spheres
Transport coefficient, x[nd2(nmkT)1I2]
Flux
Gradient
Matter
Concentration
= 3(8mn)
Momentum
Velocity
'YJ
= 5(16
Energy
Temperature
= 5 [(cv
+ 2.25k)(16m]
Chap. 2
36
Ar
N.
O.
77.7
1.29
1.26
1.33
1.33
1.60 @ 90.2 K
273.2
1.37
1.31
1.34
1.39
1.43
353.2
1.39
1.34
1.39
1.43
1.36
Temperature (K)
CR.
a Data from H. H. Landolt and R. Bornstein, Zahlenwerte und Funktionen, 6th ed. (Berlin:
Springer, 1969), Vol. II, Part 5a, pp. 3, 516.
different temperatures in Tables 2.3, 2.4, and 2.5. From Table 2.3 it is
clear that mnD/'Y) does not vary greatly. It is closer to 1.2 for monatomic
gases; deviation from the predicted ratio becomes larger as the molecules
become more complex. The data indicate that the hard-sphere model
accounts for most of the observed phenomena but that it is certainly not
precise. Table 2.4 emphasizes the reliability of the simple theory for monatomic gases (for these molecules O.4cv + 0.9k = cv ); there is very little
deviation from the predicted value of 2.5. Table 2.5 indicates that the correction for internal energy is extremely important; only if this is made do
all dilute gases behave in roughly the same fashion. Had we assumed that
kinetic and internal energies were transported in the same way, the invariant
ratio would have turned out the same as for monatomic gases, i.e., mx/ cv'Y),
a number which varies greatly. The variation of the ratio (2.43) provides
a test of the assumption that internal energy is equilibrated with each
Table 2.4. The Ratio xm/'f/cv for Monatomic Gases at Various Temperatures a
Temperature (K)
Re
Ne
Ar
90
2.44
2.46
2.49
195
2.45
2.52
2.51
2.5
273
2.45
2.50
2.48
2.50
373
2.44
2.51
2.53
2.50
2.43
2.47
2.47
2.47
491
Kr
a Data from H. H. Landolt and R. Bornstein, Zahlenwerte und Funktionen, 6th ed. (Berlin:
Springer, 1969), Vol. II, Part 4, pp. 398--451; Part 5a, p. 3; Part 5b, pp. 45-53.
Sec. 2.8
37
Temperature
+ O.9k)
H_
0_
100
2.45
(2.31)
2.37
(1.74)
200
2.50
(2.04)
2.50
(1.92)
2.28
(1.58)
2.55
(1. 78)
2.39
(1. 75)
273
2.50
(1.93)
2.53
(1.92)
2.51
(1.67)
2.60
(1.76)
2.46
(1.86)
300
2.51
(1.94)
2.58
(1.96)
2.55
(1.69)
2.65
(1.79)
2.50
(1.89)
(K)
CO_
CH.
NO
2.42
(1.80)
a Data from H. H. Landolt and R. Bornstein, Zahlenwerte und Funktionen, 6th ed. (Berlin:
Springer, 1969), Vol. II, Part 4, pp. 398-451; Part 5a, p. 3; Part 5b, pp. 45-53.
Chap. 2
38
He
20
_ _--Ne
_-----H2
10
200
400
600
T(K}
Fig. 2.7. Reciprocal of apparent value of d 2 as determined from thermal conductivity using
hard-sphere transport theory (Table 2.2). Transport coefficient data from H. H. Landolt
and R. Bornstein, Zahlenwerte und Funktionen, 6th ed. (Berlin: Springer, 1969), Vol. II,
Part 5b, pp. 45-53.
Sec. 2.8
,...... . - '
39
-------- -.---------.-
/~
------
---
_ - - - - - - Ar
. . . .-----. . ~-==--.:::::----~
CH
....- -- ---
//
/.::;.-
Ne
.-----
-.:=
--
'-C02
200
400
T(K)
600
Fig. 2.8. Reciprocal of the apparent value of d' as determined from viscosity ( - ) and selfdiffusion coefficients (- -) using hard-sphere transport theory (Table 2.2). Transport coefficient data from H. H. Landolt and R. Bornstein, Zahlenwerte und Funktionen, 6th ed.
(Berlin: Springer, 1969), Vol. III, Part 5a, pp. 3, 516.
Table 2.6. Molecular Properties of Gases and Related Transport Data Measured
at 273.2 K and 1 atm"
Gas
Viscosity,
Thermal con- Mean free
ductivity, u x 10' path, A
1) x 105
(kg m- 1 sec 1 ) (J K -1 m- 1 sec- 1 )
(nm)
Fromu
From critical
point data
168
0.218
0.218
0.250
He
1.85
Ne
2.97
4.60
120
0.258
0.258
0.224
Ar
2.11
1.63
61
0.364
0.365
0.273
H,
0.845
108
0.272
0.269
0.260
0,
1.92
2.42
62
0.360
0.358
0.272
CO,
1.36
1.48
37
0.464
0.458
0.296
CH.
1.03
3.04
47
0.414
0.405
0.299
14.3
16.7
a Data from H. H. Landolt and R. Bornstein, Zahlenwerte und Funktionen, 6th ed. (Berlin:
Springer, 1969), Vol. II, Part 5a, p. 3: Part 5b, pp. 45-53.
Chap. 2
40
where
(l6)
Sec. 2.B
+
(0 )
(e)
41
-+-__-J\
~+
(b)
(d)
Fig. 2.9. Trajectories of a particle which interacts with a center of force via a LennardJones potential for various reduced impact parameters b* = bla. As b* increases from 0,
the path changes from a head-on to a glancing collision. Plots are for different reduced
collision energies, T* = f!c~2/2E: (a) 0.8, (b) 4, (c) 20, and (d) a hard-sphere potential.
For the smaller b* the trajectories are not greatly affected by changes in T* although the
deflection angle is energy dependent. For the larger b* there are qualitative differences,
most evident in (a), where T* is the smallest. [Data from J. O. Hirschfelder, C. F. Curtiss,
and R. B. Byrd, Molecular Theory of Gases and Liquids (New York: John Wiley, 1954),
pp. 1132-1146.]
Chap. 2
42
Ne
Ar
Kr
N.
O.
CO.
CR,
Elk
(K)
35.7
a
(nm)
0.2789
35.60
a
(nm)
0.2749
124
0.3418
119.8
0.3405
190
0.361
171
0.360
91.5
0.3681
95.05
0.3698
113
0.3433
117.5
0.358
190
0.3996
189
0.4486
137
0.3882
148.2
0.3817
a Data from J. O. Hirschfelder, C. F. Curtiss, and R. B. Byrd, Molecular Theory o/Gases and Liquids (New York: John Wiley, 1954), pp. 1110-1112.
Sec. 2.8
43
5
200
I
I
I
I
u(r)/k
0
0.2
10 .4
~\
V.
"
100
(K)
log u/k 3
0.2
0.4
r(nm)
0.6
r (nm)
-100
For an improved, simpler u(r) see R. A. Aziz and H. H. Chen, J. Chern. Phys. 67, 5718
(1977).
Chap. 2
44
I lex, y I u, v) I du dv
lex, y I u, v)
(~~t (~~
(~~t (~t
r cos 0,
y = r sin 0
cos 0 -r sin 01
sin 0
rcosO
45
Problems
Problems
2.1.
What is the probability that an A atom in air at 300 K and 1 atm travels
1 [Lm without making a collision? 10 [Lm? (Use the data of Table 2.6 and
assume N2 and O. have the same effective molecular diameters.)
2.2.
Show that integration of (2.9) yields (2.10). [Choose the polar axis in the
direction of C1 so that C12 = (Ci' + co' - 2c 1c. cos 0.)1 /'.]
2.3.
Use the Jacobian method to show that the differential volume element in
spherical polar coordinates is c' sin 0 dc dO drp.
2.4.
Use the Jacobian method to prove the statement made in Section 2.4,
= dCM dc; the coordinate transformation is given by (2.15).
dC 1 dc.
2.5.
2.6.
By constructing the two-dimensional analog to (2.18) show that the collision frequency in a two-dimensional gas is
2(nRT/MA)l/'[nA d AA
+ nB d AB (1 + MA/MB)l/']
mn ). "Vt ]
mnz
bTexp [ - ( ---y;- -c;;- sin---y;-
2.8.
Demonstrate that (2.30) satisfies (2.27) and fits the boundary conditions
(2.29).
2.9.
ac
Daz
c = co,
all z,
= -J = const
'
t = 0
z = 0, all
c
where s
Co
+ -Jz
D
z/[2(Dt)l/'].
[exp( -s')
nl/'s
2
- -nl/'
~ exp(-y')dy
Chap. 2
46
2.10. At 15C and 1 atm the viscosity of NH. is 9.6 x 10- 6 kg m- ' sec- ' and the
heat capacity is 0.52 cal g-l. Estimate both D and x. The observed value
of x is 0.040 J K -1 m- 1 sec ' ; comment on any difference between your
estimate and the experimental result.
2.11. The Lennard-Jones potential can be used to compute a temperaturedependent effective hard-sphere d in the following fashion. Assume that
d is the value of r at which u(r) is the mean thermal energy, 1.5kT; this
accounts, in a rough way, for the Sutherland effect. Make this assumption
and compute d(T) for A at 100, 200, 300, 400, and 500 K. Then compare
your results with the data in Fig. 2.8. Why do you think this approach works
better at high temperatures?
2.12. At 273 K the viscosity of gaseous Br, is 1.46 x 10-' kg m- 1 sec- ' ; estimate
the molecular diameter of Br, molecules. Using the data in Table 2.6 determine the m~an free path of Br, molecules in 0, gas at 273 K and 1 atm.
2.13. The diffusion coefficient in a binary mixture is D = !(I/n""nd~B)(n""kT/2)l/'
where"" is the reduced mass of an AB pair. Estimate the diffusion coefficient
of Br, in air at 273 K and 1 atm assuming that N, and 0, have the same
diameter and a mean mass.
2.14. Assume that a tube filled with air at 273 K and 1 atm is placed in contact
with a reservoir containing Br, at its vapor pressure, 0.0844 atm. How long
will it take for Br2 to become visible at a distance 0.5 m down the tube?
This concentration is approximately equivalent to a pressure of 0.006 atm.
Use the results of Problem 2.13 to solve this problem.
2.15. The speed-dependent differential flux can be determined by extension of
the arguments that led to (2.31). It is
dJ
!me 2 dJ
no_
_
3To
2kTo
5)]
de -dT
dz
47
Generai References
nOc1
[-
2k1o
dz o
Comparison of JolKl and Joill indicates that the transport of kinetic and
internal energy involves different averaging processes. Thus x is not simply
proportional to the total molecular heat capacity, C1 + l.5k.
2.17. What are the conditions for Knudsen flow in a vessel of length 0.1 m?
Assume molecular diameters of ~0.4 nm. To what pressure does this
correspond at 298 K?
2.18. Demonstrate that (2.37) follows from (2.36).
General References
Theory of Gas-Phase Transport
L. Boltzmann, Lectures on Gas Theory, translation by S. G. Brush (Berkeley: University
of California Press, 1964).
S. Chapman and T. G. Cowling, Mathematical Theory of Non-Uniform Gases, 3rd ed. (Cambridge: The University Press, 1970), Chapters 5-8.
J. O. Hirschfelder, C. F. Curtiss, and R. B. Byrd, Molecular Theory of Gases and Liquids
(New York: John Wiley, 1954), Chapters 1, 7 and 8.
M. H. C. Knudsen, The Kinetic Theory of Gases, 3rd ed. (New York: John Wiley, 1950).
W. Kauzmann, Kinetic Theory of Gases (New York: Benjamin, 1966), Chapter 5.
J. E. Mayer and M. G. Mayer, Statistical Mechanics (New York: John Wiley, 1940),
pp. 18-30.
R. D. Present, The Kinetic Theory of Gases (New York: McGraw-Hili, 1958), Chapters 3,
4, 7, and 8.
N onequilibrium Thermodynamics
D. D. Fitts, Nonequilibrium Thermodynamics (New York: McGraw-Hili, 1962), Chapters
1-3.
S. R. de Groot and P. Mazur, Nonequilibrium Thermodynamics (Amsterdam: NorthHolland, 1962), Chapters 2--4.
I. Prigogine, Thermodynamics of Irreversible Processes, 3rd ed. (New York: Interscience,
1967), Chapters 1-5.
Electrolytic Conduction
and Diffusion
3.1. Introduction
There are a number of features which distinguish kinetic problems in
solutions from those in the gas phase. The most obvious is that the chemical
entities being studied are different in the two phases. A sodium ion in the
gas phase is a bare ion. In solution it carries with it a sheath of solvent,
i.e., it is solvated. This difference is apparent in the equilibrium properties
of solutions. In most instances there are both enthalpy changes and volume
changes upon solvation. These effects clearly indicate that the solvated
species and the pure substance differ significantly, differences which also
affect the kinetic properties.
The solvent has another important effect. Unlike the gas phase, in
which the medium is a vacuum through which molecules travel in linear
paths except for the occasional instance of interaction, collision, and
deflection, in solution the molecules of the species of interest are always
interacting with solvent. This interaction radically alters the molecular
motion. Instead of traveling in straight lines the molecules move randomly
through the maze provided by the solvent. One can describe the average
motion of molecules in solution by means of the laws of diffusion; the
motion of individual molecules is subject to such a variety of influences
that precise description is hopeless.
Solvents of high dielectric constant like water stabilize ionic species.
Thus much interest in kinetic properties in solution necessarily focuses
on the behavior of ions. This significantly complicates the discussion
49
50
Chap. 3
it
since the coulombic interaction extends over a long range. The requirement
of electro neutrality means that ions cannot diffuse through a solution
independently of one another even in very dilute solution.
Before we treat the phenomenology of chemical kinetics, assuming
we wish to discuss ionic reaction in solution, we must consider the nature
of electrical conduction in solutions. To this end we first review some properties of electrolytes and the mechanism for transport of electric current
in solution. We shall find that electrical conduction and solute diffusion are
processes that are inextricably coupled, a coupling that can be used to
relate the transport coefficient describing electrical conduction, the conductivity A, with that describing diffusion, the diffusion coefficient D.
By incorporating some ideas from the general theory of fluid flow we shall
be able to relate A (or D) to molecular properties of the solvated species.
Conductance measurements provide a way to determine D, thus establishing
limits for the reaction rate in solution. In addition, when carried out in
high electric fields, such measurements can be used to imply the existence
of ion pairs, species that are different from both solvated molecules and
solvated ions. For simplicity, and because it is the case of greatest chemical
interest, our discussion will be limited to constant temperature and constant
pressure processes.
(3.1)
Here j is the current density, or equivalently, the charge flux. The electric
field E, which is the gradient of the potential 'If', provides the driving force
for conduction. The transport coefficient a' is the specific conductivity,
expressed in ohm-I m- I . While (3.1) may appear unfamiliar, it reduces
to the common expression I = VIR in the special case of conduction in a
wire of uniform cross section (see Problem 3.1).
Sec. 3.2
51
a' - ao
(3.2)
A
Since
alC
(3.3)
(3.4)
A 0.1 M solution of NaCI is also 0.1 N. However, a 0.1 M solution of Na 2 SO. is 0.2 N;
it contains 0.2 equivalents of positive (and negative) charge per liter.
Chap. 3
52
1.5
I\OPP
x 10
(m 2 ohm-I eq-I)
HAc
0.4
0.5
AO - aC1I2
(3.7)
where AO is the equivalent conduction at infinite dilution. Here a is a speciesdependent constant. A model accounting for the concentration dependence
is discussed in Section 3.7.
From (3.6) and (3.7) it is apparent that a quantity of basic interest in
electrolytic conductance is the ion equivalent conductance at infinite
dilution. Ordinary conductance measurements cannot separate the contribution of individual ions. By designing experiments that distinguish
between ion migration toward positive and negative electrodes in a currentcarrying cell it is possible to measure A+ and L separately. A tabulation
of the results of such measurements is given in Table 3.1. (2)
(2)
Sec. 3.3
Nernst-Einstein Relation
53
Anion
in Aqueous Solution
A" x 10'
(m 2 ohm- 1 eq-l)
H+
3.498
OH-
1.978
Li+
0.3866
F-
0.544b
Na+
0.5011
CI-
0.7635
K+
0.7352
Br-
0.7820
Rb+
0.778
1-
0.769
Cs+
0.773
NO a-
0.7144
Ag+
0.6192
CIO a-
0.646
NH.+
0.734
BrOa-
0.558
!Mg2+
0.5306
10a-
0.405
!Ca2+
0.5950
Acetate-
0.409
!Sr2+
0.5946
HSO.-
0.50b
!Ba 2+
0.6364
HCO a-
0.445 b
!Cu2+
0.54
CN-
0.78 b
iZn +
0.53
!C,O!-
0.240
iNi2+
0.49 b
iSO!-
0.800
!Fe 2+
0.54b
tFe(CN)."-
1.00
!Pb2+
0.70b
!Fe(CN).4-
1.11
tFe3+
0.684b
,1.0 x 10'
(m 2 ohm- 1 eq-l)
,1.0,
tCr"+
0.67 b
tLa3+
0.695
Except where noted data from H. S. Harned and B. B. Owen, The Physical Chemistry of Electrolyte Solutions, 3rd ed. (New York: Reinhold, 1958), p. 231.
H. H. Landolt and R. Bornstein, Zahlenwerte und Funktionen, 6th ed. (Berlin: Springer, 1969),
Vol. II, Part 7, pp. 258-267.
Chap. 3
54
. + z+F
i+ =
J+ = J+(dlff)
-D+
[7
n+A+ [7
n+ - - F - 1p
(3.9)
J-
-D_ [7 n_
[7
+ -n_A_
F - 1p
(3.10)
where the sign reversal in the second term reflects the fact that a potential
gradient drives positive and negative ions in opposite directions.
To establish the relationship between D and A assume, for simplicity,
that the anionic species is essentially immobile (as would be the case were
it very large) so that J_ = O. Now assume further that the potential gradient
can be adjusted, if only in principle, to balance the concentration gradient
so that J + = O. The necessary condition, from (3.9), is
(3.11 )
With these constraints the system remains electrically neutral throughout
as there is no net ionic flow.
(3)
This formulation is based upon the simple form of Fick's law (2.23). In fact the flux
of a species depends not only on its own concentration gradient but also on the
concentration gradients of all other species in solution. A more complete theory would
have to account for such phenomena. While the precise expressions may be affected,
the general conclusions are not significantly different from those based upon (3.9)
where cross-diffusion terms are ignored. This and other coupled phenomena are discussed
in S. R. de Groot, Thermodynamics of Irreversible Processes (Amsterdam: North-HoIland, 1963).
Sec. 3.3
Nernst-Einstein Relation
55
Consider the potential energy of the cations due to the electric potential;
it is simply
14)
15)
We can now recognize why only the cation is viewed from an equilibrium standpoint.
Its electrochemical potential is fi+ = 11+" + RTlnn+ + z.FIJ! which, with (3.12), is
fi+ = Il+" + RTln n+". Since both Il+" and n+" are position independent so is fi+, as
it must be in equilibrium. The electrochemical potential for the anion is fi- = IC" +
RTln n_ - z_FIJ!. Since electroneutrality demands that z+n+ = Ln_ we find, using
(3.12), that fi- = 11_" + RT In n_" - (z+ + z_}F1J! which, since IJ! is position dependent, is not constant. Thus the anions are not at equilibrium and we see the importance
of the assumption of anionic immobility for establishing (3.13).
W. Nernst, Z. Physik. Chem. 2, 613 (1888).
Chap. 3
56
which when substituted in (3.15) leads, with (3.16), (3.17), and the observation that J = J +/v +, to alternative forms
Sec. 3.5
Stokes' Law
57
(3.20)
a result derived assuming that the fluid close to the particle is moving with
the same velocity as the particle, i.e., the fluid sticks to the particle. A great
body of experimental evidence demonstrates that (3.20) accurately accounts
for the behavior of macroscopic particles. How and why (3.20) can be used
to describe the motion of ions in a solvent is unclear. From the standpoint
of the ion the solvent is certainly not a continuous medium; it has a
molecular structure. (7) Nonetheless we shall assume that (3.20) describes
the average force acting on ions migrating in a solution. A partial justification is found in the molecular insights that will be obtained. However,
the application of (3.20) to ion migration must be considered to be a
postulate, not a consequence of fluid mechanics.
In an electric field E = -17"'1/', the force on a representative ion of
charge z is zeE; if the ion is not being accelerated this force is balanced
by friction in the same direction so that
zeE = 6Jt'Yjrv
(3.21)
The ion mobility u, defined as velocity per unit electric field strength, (8)
is therefore
(3.22)
u = vjE = zej6JtYjf
To relate u to A,o consider an ionic species with mean velocity v; the ion
(7)
(8)
It is a question of some interest whether the solvent can be presumed to stick to the
solvated ion as it diffuses through the solution. If instead it is assumed that the solvent
slips by the solute the factor 6.n in (3.20) is reduced to 4.n. This has no great qualitative
effect on the picture to be developed. For a derivation of Stokes' law see R. M. Fuoss
and F. Accascina, Electrolyte Conductance (New York: Interscience, 1959), pp. 53-59.
Theoretical studies on hard-sphere liquids indicate that Stokes' law applies if it is
assumed that solvent slips by the molecules [B. J. Alder, D. M. Gass, and T. E. Wainwright, J. Chern. Phys. 53, 3813 (1970)].
Since E, v, and j are oriented in the same direction, they are proportional to the same
unit vector n = vii v I = Ell E I = j/l j I. It is thus permissible to "divide" the vectors since their n dependence cancels.
Chap. 3
58
II
Table 3.2. Comparison of Crystal Radii, rc , with Ionic Radii in Aqueous Solution,
r A , as Estimated from (3.23)
Species
H+
Li+
Na+
K+
Rb+
Cs+
Mg'+
rA
rca
(nm)
(nm)
0.026
0.23
0.18
0.12
0.12
0.12
0.35
0.06
0.10
0.13
0.15
0.17
0.07
Species
CaH
SrH
Ba'+
OH-
ClBr1-
rA
rca
(nm)
(nm)
0.31
0.31
0.29
0.046
0.12
0.12
0.12
0.10
0.11
0.14
0.18
0.20
0.21
a L. Pauling, The Nature of the Chemical Bond, 3rd ed. (Ithaca: Cornell, 1960), p. S18.
from which, using (3.3), the ion equivalent conductance at infinite dilution
can be calculated,
;'0
(3.23)
Thus ;'0 can be used to determine ionic radii in solution. Some values of r
are given in Table 3.2; for comparison, crystal radii of the ions are also
inc1uded.(9l Except for H+ and OH-, for which r appears to be absurdly
small, the Stokes' law values are reasonable and comparable to the crystal
radii. However, the trends in both the alkali and alkaline earth ion series
are opposite. In part this may be rationalized by recognizing that a small
ion, like Li+, binds its water of solvation more strongly than a larger ion,
like Cs+; this suggests that the effective ionic radius of solvated Li+ could
be larger than that of Cs+. However, the major source of the discrepancy
is the use of Stokes' law itself. Applying a formula designed to describe
the forces on macroscopic objects in continuum fluids to a molecular
(9)
Using the slip condition (see footnote (7) in this Chapter] the values of r estimated
from (3.23) increase by 50%.
Sec. 3.5
Stokes' Law
59
(3.24)
La 3 +
1.0
-----------Ca2 +
-------____
'1 (t)
).0(t)
,,\(0) ).(0)
Fe(CN):- - - - - K+
ow
0.5
O~~----~~----~~------i-----o
50
100
150
t(Oc)
Fig. 3.2. Test of Walden's rule. The ratio [1](t) ).Ct)/1](O) ).(0)] for a series of ions as a
function of temperature.
60
Chap. 3
Table 3.3. Tests of Walden's Rule. The Product AOrJior KCI and NaCI in MethanolWater Mixtures at 25 0 ca
A1) x 10"
Methanol (wt. %)
1)
x 103
(kg m- i sec-i)
0
25
NaCI
0.8949
1.341
1.131
1.475
1.334
1.150
50
1.54
1.15
1.01
75
1.15
0.896
0.808
100
0.541
0.565
0.527
a H. S. Harned and B. B. Owen, The Physical Chemistry of Electrolyte Solutions, 3rd ed. (New
York: Reinhold, 1958), p. 706.
Table 3.4. Tests of Walden's Rule. The Product A1) for KI in Various Solvents
at 25 0 ca
Solvent
Ax102
(ohm- i m 2 eq-i)
1) x lOS
(kg m- i sec-i)
AI'j X 105
Acetonitrile
1.982
0.345
0.684
Acetone
1.855
0.316
0.586
Nitromethane
1.240
0.611
0.758
Methanol
1.148
0.546
0.627
Ethanol
0.509
1.096
0.560
Furfural
0.431
1.490
0.642
Acetophenone
0.398
1.620
0.644
Waterl'
1.5038
0.8937
1.344
S02c
2.65
0.394
1.044
a J. O'M. Bockris and A. K. N. Reddy, Modern Electrochemistry (New York: Plenum Press,
1970), Vol. 1, p. 386.
b H. S. Harned and B. B. Owen, The Physical Chemistry of Electrolyte Solutions, 3rd ed. (New
York: Reinhold, 1958), p. 697.
c At 00C.
Sec. 3.6
Mobility
61
H-O:H --..
O-H
H
I
O-H
(a)
H-O
O-H
H
1+
H-O-H
H
1
H-O
H
1
O-H
(b)
H
H-b:H
(e)
O-H
1
H
O-H
I
H
O-H
I
H
H
I
O-H
H
I
O-H
H
1
O-H
Fig. 3.3. Proposed mechanism for proton diffusion in water. (a) Proton jump. (b) Proton
jump. (c) Rotation of water molecule.
62
Chap. 3
as Bjerrum faults, (10) are of two types: a pair of oxygen atoms with two
hydrogen atoms between them or a pair of oxygen atoms with no hydrogen
atom between them. In some steps of the diffusional process such faults
must be formed as shown in Fig. 3.3c, where molecular rotation has broken
the hydrogen-bond network by creating an (O-H H-O) domain. While
no theory offault dynamics in water has been shown to correlate its electrical
properties, the Bjerrum model is known to account for the anomalous
electrical properties of ice. (11)
AO -
aCl/2
(3.26)
which can be accounted for if the effect of the ion atmosphere is taken
into consideration. At equilibrium an ion in solution is surrounded by
a spherical distribution of charge, just balancing the ionic charge itself,
as shown in Fig. 3.4a. When an electric field E is applied the ion migrates
in one direction while the ionic atmosphere, being oppositely charged,
moves in the opposite direction. Since the ion cloud is large it may be
treated as if it were a charged colloidal particle moving under the influence
of an electric field in a viscous fluid. In colloid chemistry this is known as
electrophoresis. As the ion atmosphere migrates it naturally tries to carry
its central ion with it; thus due to the electrophoretic effect there is an
effective field EE in the opposite direction to the applied field which decelerates the central ion.
In addition to the electrophoretic drag force, there is an additional
effect because a moving ion deforms its own ion atmosphere. As shown in
Fig. 3.4b an ion tends to pile up charge in front of it as it moves and tends
to leave a diminishing wake behind as the atmosphere tries to readjust
and reform the spherical distribution about the ion. The effect is to separate
charge which produces an electric field called the relaxation field E R , in the
direction of, but opposing the applied field. Considering both electrophoresis
and relaxation the local field at the ion is E - EE - ER . Since the mean
(10)
(11)
Sec. 3.7
Concentration Dependence 0/ A
63
Spherical
ion atmosphere
Oval ion
atmosphere
(b)
(a)
where UO is the mobility at infinite dilution. (12) Using the same arguments
which led to (3.23) yields an expression for Awhich accounts for the effect
of the ion atmosphere,
(3.27)
At very low electrolyte concentration the ion atmosphere is very diffuse
and both EE and ER are weak. As the concentration increases EE and ER
will also increase. Thus the ratio (EE + ER)/E increases as concentration
increases, accounting qualitatively for (3.26).
Estimates of EE and ER in dilute solution can be obtained using DebyeHuckel theory in its simplest form. Assume the electrolyte to be point
charges interacting via coulomb forces only. Ignore the molecular structure
of both solvent and ions. Since the electrophoretic effect is due to the migration of the ion cloud through the fluid, the reduction in ion mobility can be
calculated using Stokes' law (3.21)
v
E
ze
UE=-=--
6nfjr
(3.28)
where UE is the mobility of the ion atmosphere; r is the radius of the ion
atmosphere. From the Debye-Huckel theory of ion activity coefficients, (13)
(12)
(13)
Since the vectors E, EE, and ER are parallel they can be treated as scalars in division.
See footnote (8) in this chapter for discussion of a similar problem.
See, e.g., T. L. Hill, Introduction to Statistical Thermodynamics (Reading, Mass.: Addison-Wesley, 1960), p. 321ff.
Chap. 3
64
'0'
(\J
4>
'E
0)(
<
.c
o
(\J
1.31
1.27
1.23
.02
.04
.06
fC (MI/2)
r can be identified as
(3.29)
where E is the dielectric constant of the solvent and ni are the ion concentrations. Combining (3.28) and (3.29) shows immediately that UE ex: CI12.
The effect of the relaxation field is far harder to compute and we shall not
attempt it here. Qualitatively, this field is produced by distortion of the
ion atmosphere. The corresponding ionic displacements depend upon
individual ion mobilities. However, as in the problem of electrolyte diffusion
(see Section 3.4), ionic motions are coupled. As a result the precise dependence of ER upon the ion equivalent conductance is rather involved.
Onsager(14) developed a quantitative description of the relaxation field;
he demonstrated that, like EE, ER is also proportional to x and thus to C1I2.
Onsager's final expression for the equivalent conductance has the form
(3.30)
where A depends on 'Yj, E, and T while B is a function of A+, A_, E, and T.
A comparison of (3.30) with experiment for various electrolytes is shown
in Fig. 3.5. A more sensitive test is to compare the Onsager coefficient
(14)
Sec. 3.B
Wien Effects
65
Table 3.5. Observed and Calculated Values of A + EAO from (3.30) in Aqueous
Solution at 25 0 ca
Electrolyte
LiCI
NaNO.
KBr
KCNS
CsCI
MgCI 2
Ba(NO')2
K2SO.
2.56
2.60
2.78
2.42
2.40
4.56
5.08
4.43
2.30
2.35
2.54
2.46
2.54
4.60
4.76
5.04
a J. O'M. Bockris and A. K. N. Reddy, Modern Electrochemistry (New York: Plenum Press,
1970), Vol. 1, p. 437.
A + BAo with that determined experimentally from the slope of the equivalent conductance; some values are tabulated in Table 3.5. It is clear that
the Debye-Hiickel-Onsager theory accounts satisfactorily for the behavior
of A at low concentrations. As the ion concentration increases into regions
where Debye-Hiickel theory no longer accurately describes electrolyte
activity, there are also severe deviations from (3.30).
The theory of conductivity has been improved by eliminating some of
the assumptions implicit in its derivation. (10) The coupling of the relaxation
field and the electrophoretic field has been accounted for. The effect of
finite ion size and of ion-pair formation has also been taken into consideration.
(16)
R. M. Fuoss and L. Onsager, J. Phys. Chern. 66, 1722 (1962); 67, 621 (1963); 68, 1
(1964).
In nonpolar solvents Ohm's law breaks down at much lower field strengths; measurable
deviation occurs at E ~ 10' V m- 1 as (the dielectric constant) -+ 1.
66
Chap. 3
AB = (A+, B-)
K1
A+
+ B-
(3.31)
In the ion pair (A+, B-) the ions are not yet individually solvated but are
partially separated by interpolated solvent. The equilibrium constant Ko
is not field dependent. However, high fields naturally favor separation of
the ion pair; more ions are produced, i.e., KI is an increasing function of
the field. For this reason the field dependence of conductance in weak
electrolytes is also known as the dissociation field effect. Arguing from (3.31),
the apparent equilibrium constant is
[A +][B-]
[ABlto t
(3.32)
(18)
(19)
(20)
Problems
67
Problems
3.1.
3.2.
A+
3.3.
(21)
(22)
There is an enormous literature on electrode kinetics; see P. Delahay, Double Layer and
Electrode Kinetics (New York: Interscience, 1965).
D. Berg and A. Patterson, Jr., J. Am. Chern. Soc. 75, 5197 (1953); K. F. Wissbrun, D.
M. French, and A. Patterson, Jr., J. Phys. Chern. 58, 693 (1954).
Chap. 3
68
3.4.
3.S.
3.6.
Consider a membrane of thickness d across which there is a potential difference LltP. Show that the current density across the membrane is
j
(j
LltP
G LltP
where G is the conductivity per unit area of membrane surface. Now show
that
where C* is the concentration of charge carriers in the membrane.
3.7.
Use the data of Table 3.4 to estimate the ionic radius of K + in the various
solvents listed. You may assume, at this level of approximation, that K +
and I-have the same ionic radii. Can you suggest any explanation for the
changes observed?
3.8.
3.9.
For 1-1 electrolytes in water solution at 2SoC the values of A and B in the
Onsager equation are A = 1.90 X 10-4 ohm- 1 m' eq-1 (m3 eq-1)1/2 and
B = 7.24 X 10- 3 (m3 eq-1 )1/2. Using these values and data from Table 3.1
calculate:
(a) the specific conductivity of a 0.1 M solution of KCI; the experimental
value is 1.2896 ohm- 1 m- 1;
(b) the specific conductivity ofa 0.1 M solution of HAc (KA = 1.8 X 10- 5 M).
3.10. For a binary electrolyte the expressions for A and B in the Onsager equation
are
A = 9.218 x 1O- 6 (z+ + L)3!2/,YJ(T)l!2
3.13
1O- 4 (z+
+ L)1!'w/(T)3/2
General References
69
Z+L(A+
+ ql/2)
By using mixed solvents the dielectric constant of the medium can be varied
continuously. The values of E in H20-CHaOH mixtures at 25C are
CHaOH (wt. %):
E:
25
50
75
100
78.54
67.8
56.3
44.9
32.66
Using these data and those of Table 3.3 calculate A for lO- a M and 0.1 M
solutions of KCl for the various solvents. Do you think the estimates of A
at the higher concentrations are reliable?
3.11. Show that (3.32) follows from (3.31).
General References
J. OM. Bockris and A. K. N. Reddy, Modern Electrochemistry (New York: Plenum Press,
4
Determination of Rate Laws
4.1. Introduction
Altering the constraints on a chemical system poses two distinct
questions. What is the new equilibrium configuration? How rapidly does
the system approach this new state? The first is a problem of applied
thermodynamics; the second is the central problem of chemical kinetics.
The transport phenomena discussed in Chapters 2 and 3 are the simplest
examples of kinetic phenomena, if not the most familiar. For the cases
considered, a single macroscopic property (number, charge, momentum,
or energy density) was displaced from its equilibrium value. This perturbation caused a flux in the opposite direction, proportional to the displacement.
The proportionality constant is the transport coefficient. For simple gaseous
systems the relations between displacement and flux and the transport
coefficient were determined using a hard-sphere model to describe molecular
interaction.
Reacting systems are much more complex. Corresponding to the linear
phenomenological transport equations (Fick's, Fourier's, Ohm's, or Poiseuille's laws) are the nonlinear rate laws for chemical reactions. The
rate constant is the analog of the transport coefficient. The goal of theory
is to develop a molecular model which accounts for the rate law and which
provides an expression for the rate constant. These problems are, in many
ways, still poorly understood; a unified molecular theory such as was
developed for treating transport in gases does not yet exist. Instead, reaction
mechanisms, which are proposed as models for the underlying molecular
events, are constructed to be consistent with the observed rate laws. Then
very approximate theories for the rate constants are formulated in terms of
71
Chap. 4
72
(4.3)
Sec. 4.3
73
2Br
+ H2 --' HBr + H
H + Br2 --' HBr + Br
H + HBr --' H2 + Br
Br
(equil. dissoc.)
(slow)
(fast)
(4.5)
(fast)
was finally shown to be consistent with (4.3). Later research suggested that
(4.2) and (4.4) were deceptively simple; the first step in the formation of
HI is very likely the dissociation of the halogen. The subsequent steps are
different from (4.5) and there are still numerous unanswered questions.
The fact that seemingly simple kinetics like those in the formation
of HI could sti\l be an active subject for controversy for three-quarters of
a century is indicative of a major problem in mechanistic assignment.
It is impossible to prove a mechanism correct. Rather one shows that various
plausible mechanisms are inconsistent with the data. By amassing sufficient
chemical evidence one hopes to show that all but one chemically reasonable
mechanism is impossible. However, this can never really be done since
there are always more complex mechanisms which can account for the
data. Thus, even given compelling evidence, there remains a chance that
further experimentation will require revising the model which we call a
mechanism.
We use the notation A + B = C to describe the stoichiometry of a reaction. The notation A + B ~ C is used to mean a specific molecular encounter.
74
Chap. 4
Rieq)
kiT,p)
RrCeq)
kr(T,p)
n
n
react.
prod.
[Yd;;i
(4.9)
[Yi]~~
where the k's are the rate constants.(5) Since the two rates are equal at
(0)
(5)
For a discussion of progress variables and chemical equilibria see, e.g., F. T. Wall,
Chemical Thermodynamics, 3rd ed. (San Francisco: Freeman, 1974), Chapter 10.
The pressure dependence of k in single-step reactions is usually small; this problem is
treated by C. A. Eckart, Ann. Rev. Phys. Chem. 23, 239 (1972). Gas-phase unimolecular
reaction, for which k is a sensitive function of p, is not a single-step process (Section
5.4),
Sec. 4.3
75
TI ai
v
l
= kf(T,p)
k (T)
r ,p
TI Yi'
(4.11 )
The ratio of forward and reverse rate constants can be identified as the
Yi Vi = 1, a condition satisfied in ideal
equilibrium constant whenever
systems such as dilute gases. In addition, by redefining the standard state
of the reacting species, the constraint may be satisfied if the Yi do not
vary significantly with reactant concentration (see Problem 4.2). It is
clearly to the kineticist's advantage to study systems under conditions
where the activity coefficient product is unity. The activity correction
requires that considerable care be exercised when comparing kinetic parameters determined under grossly different conditions.
Assuming that away from equilibrium the reaction rates of the forward
and reverse steps are still given by (4.9),(7) we expect that R j *- Rr and that
the overall rate of reaction is
(4.12)
which is precisely of the form (4.2) as long as kr or [HI] are small. From
(4.11) we note that if Keq varies with temperature or pressure one (or both)
of the rate constants must do so too. Furthermore, the activity dependence
indicates that the rate constants may be concentration dependent. The
(61
(71
For a discussion of activity, activity coefficient, and their relationship to the equiiibrium
constant see F. T. Wall, Chemical Thermodynamics, 3rd ed. (San Francisco: Freeman,
1974), pp. 365-366, 419--420.
This is equivalent to assuming that the collision frequency (2.18) is unaltered in a
system displaced from equilibrium.
Chap. 4
76
react.
-viYi ~
prod.
ViYi
where /1i is the chemical potential. (8) The condition for chemical equilibrium
is that A be zero; the chemical transport equation is thus
~ = LA
(4.13)
(9)
Rieq)jRT = Rr(eq)jRT
The affinity is discussed by F. T. Wall, Chemical Thermodynamics, 3rd ed. (San Francisco: Freeman, 1974), Chapter 10.
For a simplified demonstration of this limited equivalence see A. Katchalsky and P. F.
Curran, Nonequilibrium Thermodynamics in Biophysics (Cambridge, Mass.: Harvard
University Press, 1965), pp. 91-97.
Sec. 4.4
77
+ CIO- =
10-
+ CI-
+ Br
->.
HBr
+ Br
(4.15)
Chap. 4
78
+H0
CO (aq) + H 0::;;:::'::: H+ + HCO aH+ + HCO a-::;;:::,::: H CO a
H 2CO a ::;;:::,::: CO 2 (aq)
2
(4.17)
of the cycle A ---' B ---' C ---' A even if the forward and reverse steps in
each elementary step did not proceed at the same rate. An expression
analogous to (4.16) could be derived for an equilibrium constant even
though (4.11) might not be true for the individual steps. It is precisely
such a possibility that is forbidden by microscopic reversibility.(ll)
Newtonian mechanics and quantum mechanics have in common the
property of time-reversal invariance. The Newtonian equations of motion
for a system of N interacting particles are
for all x, y,
Z,
i = 1, ... , N
(4.18)
Changing t ---' -t does not affect the accelerations Xi; on the other hand the
velocities Xi do change sign. As a consequence, for every solution of (4.18)
there is an identical one in which all particles have their velocities
reversed.o 2 ) For classical point particles the dynamics of a colliding pair
The following treatment is adapted from N. Davidson, Statistical Mechanics (New
York: McGraw-Hili, 1962), pp. 230-235.
(12) The dynamics are more complicated when magnetic fields are present. Since the field
couples with the velocity by means of terms like v x B it is also necessary to reverse
the field direction when the velocities are reversed. In this way time-reversal invariance
can be generalized.
(11)
Sec. 4.5
79
Since neither collision nor time reversal affects energy, the states 1,2,2-,
and 1- are isoenergetic and therefore equally probable in an equilibrium
system. As a consequence the number of pairs making the transition 1 --' 2
exactly equals the number which go from 2- --' 1-. This is the quantum
version of detailed balance. (14)
We can now show that each reaction in the system (4.17) equilibrates
individually. On the molecular level in an equilibrium system the number
of A --' B transitions, each molecule being in a specified state, is exactly
equal to the number of back transitions for the time-reversed molecules.
Since a molecular state and its time-reversed partner are chemically indistinguishable (the two states are degenerate) the overall effect, when averaged over a Boltzmann distribution of energies, is that the rate of forward
and backward reaction in the individual steps of (4.17) must be equal at
equilibrium. In addition detailed balance implies a similar dynamic equilibrium among molecules undergoing transitions between any subset of
states.
(13)
(14)
For particles which interact via spherical potentials a more restricted principle can be
demonstrated. Velocity reversal is not required for equivalence of transition probability
so that the transition (c 1 ', co', b') ~ (c1 , c2 , b) and its inverse (c1 , c2 , b) ~ (c1 ', co', b')
are equally probable, a result demonstrated in the derivation of the Maxwell distribution, Section 1.2.
Microscopic reversibility is the transition probability constraint. Detailed balance is
its consequence in an equilibrium system. See B. W. Morrissey, J. Chern. Ed. 52, 296
(1975).
Chap. 4
80
Sec. 4.6
81
(6) conductance;
(7) optical rotation;
(8) concentration (by removmg an aliquot, quenching the reaction,
and titrating);
(9) vapor chromatography;
(10) mass spectrometry;
(11) electron paramagnetic resonance spectroscopy.
It is not always simple to determine the concentrations from the measured
II Ei(Y)Ci(t)
(4.19)
where I is the path length, Ei(Y) the extinction coefficient of species i, and
Ci(t) its concentration. If only one species absorbs, A is directly proportional to the concentration of that species. If more than one species absorbs,
measurements are required over a range of frequencies to determine the
Ci(t). Similar difficulties exist no matter what monitoring procedure is
used. The general time-dependent response of an evolving system, R(t), is
R(t)
aiCi(t)
(4.20)
where the ai are determined by the properties of species i and by the measurement procedure used(15) (Problems 4.21 and 4.22).
In all cases the time required for a measurement (the response time
of a particular technique) must be short compared with the time during
which significant composition changes occur. For example, method 8 is
well suited to reactions with half-times on the order of hours but requires
great ingenuity when applied to reactions with short half-times. (16) The
practical limits of various techniques are summarized in Fig. 4.1.
For reactions which are essentially complete between periods ranging
from 1 msec to 1 sec rapid mixing techniques must be used. Two such
I thank Professor L. Parkhurst for pointing out to me the practical significance of this
problem.
(16) Quenching, by spraying a reacting mixture into a cold liquid or onto a cold plate,
can be completed within milliseconds. Combining this approach with flow methods allows method 8 to be applied to reactions with half-times ;s 100 msec. See B. Chance,
Q. H. Gibson, R. H. Eisenhardt, and K. K. Lonberg-Holm, eds., Rapid Mixing and
Sampling Techniques in Biochemistry (New York: Academic Press, 1966), Chapter 6.
(15)
82
Chap. 4
-flow-
10 6
__ electric field - +
p_ jump _ _ _
pulse
T - j u m p - - - - - -...
--+.
sound
10- 6
10 0
tube
- - flash --...;.......
- --
proton
_ _nmr
_
'~
10-12 1
shock.
classical --+
I limit
of
I chemistry
II
t(sec)
pulsed laser
flash
~
1
1
I
esr
_ _fluorescence
_ _ _- J '
~y~
competition with
Fig. 4.1. Practical range of applicability of various methods for measuring the rates of
chemical reaction.
procedures are the stopped flow and the continuous flow methods. In
stopped flow(17) the reactants are driven into a small reaction chamber in
such a way that mixing is essentially complete within 0.1 msec. The reaction
is then monitored automatically using sensitive recording techniques
synchronized with the arrest offluidflow. In continuous flow(18) experiments
reactants continuously enter a reaction chamber, mix, and are forced into
a flow tube where the reaction occurs. Depending upon the rate of flow,
each point in the tube corresponds to a different time after the onset of
reaction. By monitoring at different points the degree of reaction at different times can be determined.
When the reaction time is less than 1 msec, flow methods are too slow
to insure homogeneity. Shock tube(19) methods are useful for studying
high-temperature gas-phase reactions which take place in time spans of
1 msec to 1 [lsec. Here a driver gas (usually He or H2 at a few atmospheres
pressure) is constrained by a diaphragm at one end of a long tube; the
remainder of the tube is filled with reactant mixture at low pressure (typically
10-3 atm). When the diaphragm is ruptured the driver gas expands abruptly,
(17)
(18)
(19)
The basic idea for stopped flow was introduced by B. Chance, J. Franklin Inst. 229,
455, 613, 737 (1940). For a modern apparatus see Q. H. Gibson and L. Milnes, Biochem. J. 91, 161 (1964).
This method was pioneered by Rutherford for studying gas-phase reactions [E. Rutherford, Phil. Mag. 44, 422 (1897)]. The mixing methods that are now in common use
were designed by H. Hartridge and F. J. W. Roughton, Proc. Roy. Soc. (London) AI04,
376 (1923), and by G. A. Millikin, Proc. Roy. Soc. (London) A155, 277 (1936).
Shock tubes were invented by M. P. Vieille, Compt. Rend. 129, 1228 (1889). For a discussion of modern design see H. T. Nagamatsu, Fundamental Data Obtained from Shock
Tube Experiments, R. Ferri, ed. (New York: Pergamon, 1961), pp. 86-136.
Sec. 4.6
83
producing a shock wave which travels down the tube at supersonic speed.
The shock front is very sharp. As it passes any point in the reactant mixture
the reactant gases are compressed, heating them to temperatures "-' 103-10 4
K; this compression, which takes "-' 1 [isec, provides one limit for the
technique's applicability. Reaction begins as the gases are heated with
passage of the shock front. To study the system one relies on the fact that
the driver gas lags behind the shock front, taking a few milliseconds to flow
down the tube. An observation point is established near the end of the tube.
The first changes to be seen take place when the shock front passes. The
gas behind the front has been reacting from the instant the shock wave
passed by. It is pushed past the monitoring station by the driver gas. The
later the reactants pass the point, the longer they will have had to react.
Thus a time profile can be established.
Shock tubes are of limited utility. A more general approach to the
study of reactions which are complete in the range I msec-I nsec is to use
fast reaction methods. (20) An equilibrium system is perturbed by an external
stimulus applied for a very short time (always less than the half-time for
reestablishing equilibrium). A common approach is to effect a temperature
jump in the system by a brief burst of heating. If the equilibrium is temperature sensitive the concentration of reactants must readjust; by synchronizing an automatic recording technique with the onset or termination
of the heating pulse the relaxation to the new equilibrium state can be
followed. There are many other stimuli that can be used to perturb the
system. These include dilation (pressure jump), electric field (Wien effect),
etc. Any method that can perturb the system very rapidly is potentially
useful for such an experiment.
Perturbation methods displace a chemical equilibrium but no new
species are formed. In flash photolysis novel, unstable chemical species are
produced by irradiation. Their reactions with the species initially present
are the domain of photochemistry (treated in Chapter 6). Such reactions
(20)
The original fast reaction techniques were based upon flash photolysis [G. Porter,
Proc. Roy. Soc. (London) A200, 284 (1950)] and the Wien effect [M. Eigen and J.
Schoen, Z. Elektrochem. 59, 483 (1955)]. An enormous literature exists on the subject.
For recent discussions of some techniques see literature on: (1) temperature jump,
G. W. Hoffman, Rev. Sci. Instr. 42, 1643 (1971), and D. H. Turner, G. W. Flynn, N.
Sutin, and J. V. Beitz, J. Am. Chern. Soc. 94, 1554 (1972); (2) pressure jump, A. Jost,
Ber. Bunsenges. Phys. Chern. 70, 1057 (1966); (3) electric field methods, L. C. M. de
Maeyer, in Methods in Enzymology, Vol. 16, K. Kustin, ed. (New York: Academic
Press, 1969), p. 80; (4) ultrasonic methods, K. G. Plass, Acoustica 19, 236 (1967/68);
(5) flash photolysis, G. Porter and M. R. Topp, Proc. Roy. Soc. (London) A315, 163
(1970).
Chap. 4
84
~========'~~~--~EJE
I
~'
,00
(22)
For an introduction to these methods see D. N. Hague, Fast Reactions (London: WileyInterscience, 1971), pp. 47-75. Photochemical competition (quenching) is discussed in
Chapter 6.
G. Porter and M. R. Topp, Proc. Roy. Soc. (London) A315, 163 (1970).
Sec. 4.7
85
is monitored is the lag between the times when the initiating flash and the
scintillation radiation pass through the reactant vessel. The time lag is
specified by the paths of the light beams and is, therefore, fixed by the
geometry of the optical system. By varying the position of the mirror that
sends light to the scintillation solution, different time lags are produced
(0.3 m == 1 nsec). The technique has only one serious limitation: pulsed
lasers cannot produce exciting radiation at every frequency of interest.
Thus not all systems are amenable to study. Still faster reactions have been
studied using a variation of the pulsed-laser technique; the picosecond
range is now accessible. (23)
dt
-k[A]l.O
(4.21 )
dt
= _ d[B] = -k [A]1.0
dt
+ L[Bp.o
(4.22)
86
Chap. 4
[AJ
Ao - $,
[BJ
Bo
+$
Here the time dependence is also exponential but the decay constant is
the sum of the forward and reverse rate constants. Thus, for exponential
relaxation, either of the rate laws, (4.21) or (4.22), is consistent with a
concentration-independent decay constant. The individual rate constants
may be evaluated if the equilibrium constant is also known. Ignoring
activity effects, L = k +K;;J and we find
(4.23)
where x is the decay constant. If Keq ~ I, k+ is essentially equal to the
decay constant; the reverse step in (4.22) is unimportant and the two rate
laws are kinetically indistinguishable.
First-order processes are very common. The most familiar is radioactive decay, e.g., 14e ---->. 14N + e, the nuclear reaction used in carbon
dating for which k = 1.24 X 10-4 yr- 1 An interesting chemical example
is the gas-phase decomposition of N 20 5 which obeys the stoichiometry
with k = 3.38 X 10-5 sec- 1 at 298 K. This suggests a first step of fragmentation of an N 20 5 molecule; it provides no information about the
fragments produced or the subsequent steps in the reaction. To detail
the further course of reaction more experiments are needed. Examples
of the reversible first-order rate law (4.21) are found in many reactions
involving rearrangement, isomerization, or conformational change. An
interesting case is the chair 1 ---->. chair 2 rearrangement in cyclohexane shown
in Fig. 4.3. The rate law is
d[chair ]1
dt
-k[chair]l
+ k[chair12
Sec. 4.8
Isolation Methods
87
~2~
Fig. 4.3. Chair! ~ chair. interconversion in cyc10hexane showing exchange of axial and equatorial hydrogen atoms.
(2)
since the specific rate constants for both forward and backward steps must
be equal; at 206 K the rate constant is 52.5 sec- 1 .(25)
+ a + ao
1;
where the coefficients ai depend upon the initial concentrations and stoichiometric coefficients of the reacting species and upon the rate constants
in a variety of complicated ways. Integration of the rate expression always
yields a result that involves three parameters,
[A]
-A
=1+
o
ae(l - e- At )
At
1
ee-
(4.24)
where a, e, and A are determined by the aj. (26) To derive and use such expressions for eliciting a rate law is a time-consuming, demanding, and
error-prone task. (27)
F. R. Jensen, D. S. Noyce, C. H. Sederholm, and A. J. Berlin, J. Amer. Chem. Soc. 84,
386 (1962). The experiment is particularly interesting. Reactants and products are chemically indistinguishable but it is still possible to investigate their rate of interconversion.
The reaction rate is deduced from nuclear magnetic resonance line-width measurements.
(26) The same form is deduced if the rate laws of Table 4.1 describe nonelementary systems
for which autocatalysis (A or B as products) or autoinhibition (C or D as reactants)
occur. However, in some such instances A may be negative or imaginary; as a consequence unchecked growth or oscillatory behavior may be predicted. In Chapter 7 we
shall see that such behavior cannot describe a reacting system at all times although
it may be approximately correct for some periods in systems far from equilibrium.
(27) For a compendium of integrated rate laws see A. A. Frost and R. G. Pearson, Kinetics and Mechanism, 2nd ed. (New York: John Wiley, 1961), Chapters 2 and 8.
('5)
88
Chap. 4
-d[A]
dt
Rather than determining a rate law under conditions where all concentrations vary simultaneously, it is easier to isolate one species and hold
the concentration of the others constant. Consider a reaction which goes
to completion and follows the rate law
(4.25)
where a, b, and c are the unknown exponents. If the reaction mixture is
flooded with Band C so that during reaction their concentration is invariant,
the apparent rate law is
(4.26)
which may be easily integrated. If a = 1.0, the kinetics are pseudo-firstorder; the dependence of " on the initial concentrations Bo and Co then
establishes the exponents band c. When a -::/= 1.0, decay is not exponential;
the integrated form of (4.26) is
[Ap-a
(a - l)"t
+ Aol-a
(4.27)
d~~]
-k1[A]a[B]b _ k 2 [A]a'[C]c
(4.28)
d~]
-k+[A]a[B]b
+ L[C]C[D]d
(4.29)
Sec. 4.9
a
I
89
Relaxation Methods
= I. 5
=2
= I. 25
J[A)4
f[A]4
2
2
2
t
234
Fig. 4.4. Plots of [A]l-a vs. time for various choices of a in (4.27). A linear relation is
assumed when a = 1.5.
-k C c[A]a l
2
and normal isolation experiments can be carried out. The rate law (4.29)
includes a step in which A is produced. If initially there is a large excess
of B the reaction, in its early stages, appears to be of order a; as reaction
progresses it is affected by the reverse step and isolation conditions no
longer apply. To maintain isolation one of the reaction products must be
removed, as is possible if it precipitates or is volatile.
A practical indication that isolation conditions have not been established is when the trial and error method for determining the exponent a
outlined in Fig. 4.4 fails. Then concentrations other than that of A vary
significantly during the course of reaction. If no changes in the experimental
conditions are adequate to circumvent this problem a different approach
must be used.
+ Aeq)/VA =;,
(-[B]
+ Beq)/VB =;,
etc.
(4.30)
90
Chap. 4
~;
+k_(cvcDeq
=
+ dVDCeq)C~~lD~;;l]
(4.31)
-x~
The first term in (4.31) vanishes since, at equilibrium, the rate of reaction
is zero. When integrated (4.31) is
(4.32)
a result only valid when ~o is small. The decay constant x is the sum of two
terms, characteristic of the forward and of the reverse step in the reaction.
In general we can write
(4.33)
where Xi is the contribution to the decay constant from the forward step
in the reaction, etc. In Table 4.2 expressions for xi are given for a number
of rate laws; with these results it is possible to construct the decay constant
that describes any relaxation experiment close to equilibrium. The relation
between the order of a particular step in the reaction and the concentration
dependence of the corresponding term in x is particularly simple. For a
first-order step x is concentration independent, for a second-order step x
Table 4.2. Contribution of Forward Step in Reaction to Overall Decay Constant
for Different Rate Laws
Forward rate expression
Xf
(defined in text)
k+[A]l.O
k+
k+[AP' o
4k+Aeq
k+ [AP' o[B]l.O
k+(Aeq
k+[A]3.o
9k+A~q
k+ [A]"'o [B]l.O
k+(4AeqBeq
+ Beq)
+ A~q)
k+(AeqBeq + AeqCeq + BeqCeq)
Sec. 4.9
Relaxation Methods
91
for the rate law (4.29). Unless [k+(T)jL(T)] = [k+(To)jL(To)] the rates
of forward and reverse steps are no longer equal and the system must
readjust to a new equilibrium state. Analysis in terms of the progress
variable leads to the previous result, (4.32). Naturally there must be a
(28)
(29)
(30)
Fora general treatment of the concentration dependence of decay constants for a system
of sequential reactions see G. w. Castellan, Ber. Buns. Gesell. 67, 898 (1963).
A useful method for separating many overlapping exponentials is given by R. D. Dyson
and I. Isenberg, Biochemistry 10, 3233 (1971). Where only two processes overlap a
simple subtraction procedure can be used; see D. S. Honig and K. Kustin, Inorg. Chern.
11, 65 (1972) for an example.
M. Eigen, Disc. Faraday Soc. 17, 194 (1954); R. G. W. Norrish and G. Porter, Disc.
Faraday Soc. 17, 40 (1954).
92
Chap. 4
hO
'c'"
0
~N 1
Pyridine-4-aldehyde
(P)
+
Piperazine
(A)
Carbinolamine
(C)
(32)
(33)
(34)
(35)
VIa
parallel
The general constraints which limit the application of perturbation methods are derived
by M. Eigen and L. de Maeyer, in Techniques of Organic Chemistry, 2nd ed., A.
Weissberger, ed. (New York: Interscience, 1963), Vol. 8, Part 2, pp. 929-941.
P. George and G. I. H. Hanina, Disc. Faraday Soc. 20, 216 (1955).
D. E. Goldsack, W. S. Eberlein, and R. A. Alberty, J. Bioi. Chern. 240, 2312 (1965).
For analysis of the consequences of different heating pulses see M. Eigen and L. de
Maeyer, in Techniques of Organic Chemistry, 2nd ed., A. Weissberger, ed. (New York:
Interscience, 1963), Vol. 8, Part 2, pp. 917-928.
H. Diebler and R. N. F. Thornley, J. Am. Chern. Soc. 95, 896 (1971).
Sec. 4.10
93
where AH + and CH + are the protonated forms of piperazine and carbinolamine. Furthermore, the possible mechanism
PH++A~CH+
Chap. 4
94
law can be tested for consistency with the kinetic measurements and the
rate constants can be determined.
In general the kineticist is faced with the problem of deducing rate
laws that describe consecutive reactions involving many steps. The corresponding rate equations are coupled, and, in most cases, the functions
that are to be fit cannot be determined using linear least-squares analysis.
A vast literature of nonlinear function-fitting methods exists to treat these
problems. (37)
+ Fe (III) =
Eu(III)
+ Fe(lI)
dt
= k'[Eu(U)][Fe(III)]
(4.34)
While the fit is gratifying, it is hardly conclusive since many other functions
are qualitatively similar, e.g.,
k' = a/(b
+ [H+])
(4.36)
(38)
For a brief discussion, with examples from enzyme kinetics, see J. T. Wong, Kinetics
of Enzyme Mechanisms (London: Academic Press, 1975), Chapter 11. A more extensive
development is given by Y. Bard, Nonlinear Parameter Estimation (New York: Academic Press, 1974).
D. W. Carlyle and J. H. Espenson, J. Am. Chem. Soc. 90, 2272 (1968).
Sec. 4.10
95
Table 4.3. Apparent Binary Rate Constant, k',/or the Reaction of Eu(JI) and Fe(JII)
as a Function of [H+] and Temperature a
T= 1.4C
k' x 10-4
[H+]
(M)
(M-I sec-I)
0.031
0.04
0.05
0.1
0.2
0.5
0.879
0.939
8.14
6.92
4.92
2.70
1.33
0.729
0.616
0.613
T= 15.8C
k' x 10-'
[H+]
(M)
(M-I sec-I)
0.031
0.036
0.05
0,07
0.1
0.13
0.2
0.876
1.00
22.9
17.9
13.4
8.54
6.86
5.91
3.30
1.51
1.25
T= 25.0oC
k'
10-4
[H+]
(M)
(M-I sec-I)
0.03
0.04
0.05
0.0667
0.08
0.111
0.167
0.2
0.4
0.953
51.7
38.9
28.8
24.6
18.1
13.3
9.14
7.65
4.08
2.11
Fig. 4.5. Temperature and hydrogen-ion dependence of apparent binary rate constant in
Fe(III) + Eu(II) system. 0, 25.0C. 0, 14.8C. /::., 1.4C. [Adapted from D. W. Carlyle
and J. H. Espenson, J. Am. Chern. Soc. 90, 2272 (1968).]
96
Chap. 4
kl
+ L~
M
k_l
. L(H 20)
+ H 0,
2
k.
+ L~M
2L + H 0,
k_.
KI = 1893 M-I
(4.38)
where the equilibrium constants were determined at 25C. (41) The approach
to equilibrium was followed by a spectrophotometric technique which was
sensitive to the concentration of the dirhodium complex M 2H 20. The
decay constant was determined at 25C for a variety of total concentrations
of both M and L, CM, and CL; the data are given in Table 4.4.
(39)
(40)
(41)
When (4.36) is assumed to represent k, the correlation coefficient is 0.964, a result that
would occur by chance only once in 104 ; of course, the hypothesis (4.35) correlates
substantially better. The inappropriateness of (4.36) is only demonstrated when k- 1
is plotted vs. [H+] (see Problem 4.10).
K. Das, E. L. Simmons, and J. L. Bear, Inorg. Chern. 16, 1268 (1977).
L. Rainen, R. A. Howard, A. P. Kimball, and J. L. Bear, Inorg. Chern. 14,2752 (1975).
Sec. 4.10
97
CL
xx 10-3
(mM)
(mM)
(sec-I)
[M]
(mM)
[L]
(mM)
1.94
1.96
5.88
0.79
0.67
1.92
2.91
6.25
0.50
1.21
1.92
3.38
7.14
0.41
1.51
1.91
3.85
7.69
0.33
1.84
1.90
4.31
8.70
0.27
2.19
1.89
4.76
9.09
0.23
2.54
= Kl
(4.40)
+ [ML] + [ML
[L] + [ML] + 2[ML
[M]
2]
(4.41 )
2]
The values of [M] and [L] are also included in Table 4.4. To test whether
(4.39) adequately accounts for the data, X is plotted against [M] + [L] +
IIKI in Fig. 4.6. According to (4.39) we expect a straight line passing through
the origin in good agreement with the data. If it is assumed that the line
passes through the origin, linear least-squares analysis (see Appendix)
98
Chap. 4
Or-----,-----~------.-----,
OL-----~I----~2------3~--~4
Sec. 4.11
99
ken
Ea
RP d(ln k)
dT
-R d(ln k)
d(lIT)
R()
+ mRT
(4.44)
Ea(high T) - Ea(low T)
= ---''-;;R:-::(~hi;-g7""h-';:;T;:;----:1'=-0'':''''w--;T:;:;):---=--
(4.45)
Chap. 4
100
9r-r--------.------~r-------._,
..
(/)
::2:
.::.t.
cii
E
.002
.003
.004
.005
rl(K)
Fig. 4.7. Arrhenius plot of the rate constant for the reaction Clep) + CH 4 ~ HCI + CH.
Note the definite curvature [D. A. Whytock, J. H. Lee, J. V. Michael, W. A. Payne, and
L. J. Stief, J. Chern. Phys. 66, 2690 (1977)].
Appendix
101
y=kx
need not be zero. If the functional form (4A.l) is correct and all experimental
errors are random the best estimate of k is found by minimizing the quantity
Li di 2 = I
(45)
(46)
Yi 2 -
2k I
XiYi
+ k2I
Xi 2
(4A.3)
See F. T. Wall, Chemical Thermodynamics, 3rd ed. (San Francisco: Freeman, 1974),
Chapter 3.
For an introduction to data analysis see H. D. Young, Statistical Treatment of Experimental Data (New York: McGraw-HilI, 1962), in particular, pp. 101-132. A comprehensive treatment is given by N. R. Draper and H. Smith, Applied Regression Analysis
(New York: John Wiley, 1966), Chapter 10.
Chap. 4
102
and we obtain
(4A.4)
kx
+q
(4A.5)
(4A.6)
where
(4A.7)
When r = 1 there is perfect correlation, i.e., (4A.5) is exact for each datum;
when r = 0 there is no correlation. Since 1 > r > 0 in any series of experiments we may refer to tables of correlation coefficients(48) for the
answer to this question: What is the probability that the same value of r
would be found by chance if the variables were unrelated? Finally the
standard deviation in k and q can be calculated (49):
(h = a(NjiJx)1I2,
a
(47)
(48)
(49)
aq
--,--
a(L Xi 2jiJX)1I2
(4A.9)
See H. D. Young, Statistical Treatment of Experimental Data (New York: McGrawHill, 1962) for details.
See H. D. Young, Statistical Treatment of Experimental Data (New York: McGrawHill, 1962), p. 164 for such a table.
If q = 0 then correlation is via (4A.l). In this case Gq == 0, r = ~ XiY.J(~ Xi2~ y.j2)1I2,
and Gk = ~(yj - kXj)2/~ Xj2]1/2.
103
Problems
Problems
4.1.
Recent research [J. H. Sullivan, J. Chem. Phys. 46, 73 (1967); 39, 3001 (1963)]
has indicated that the H. + I. reaction may proceed by one of the two following sets of elementary steps:
I.
(i)
21
(equilibrium dissociation)
H. + 21~2HI
(ii) H. + 1 = H.I
H.I
(equilibrium addition)
k'
+I~2HI
(a) Show that either pathway accounts for the overall stoichiometry.
(b) Write the rate law for each pathway and show that each is consistent
with Bodenstein's experimental results (4.2).
(c) Relate the overall equilibrium constant to the rate constants of the
individual steps. You may need to include some other reactions.
4.2.
where ili =
4.3.
fli*
IT exp( -vdidRT)
+ RT In yj(l).
Co(NH.),(H.O)3+ + HgBr+
has the approximate ionic strength dependence, log k f = log kl+4.1(I 1 /').
[J. N. Bronsted and R. Livingston, J. Am. Chem. Soc. 49, 435 (1927)].
Use Debye-HiickeI theory to estimate the activities and thus obtain the
ionic strength dependence of kr .
4.4.
104
Chap. 4
Table 4.5
t (sec)
0.0450a
0.0899
0.135
0.180
0.5
0.45
0.55
0.25
0.15
0.15
0.4
1.15
0.55
0.35
0.25
0.35
1.80
0.85
0.55
0.45
0.3
2.50
1.20
0.75
0.60
0.25
3.20
1.55
0.95
0.80
0.2
4.05
1.95
1.20
1.00
(b) From the decay constants determine the dependence of the rate law
on [H 20,].
4.5.
Show that T (temperature) jump is a useful kinetic probe only if ;JHrxn =F- O.
Show that P (pressure) jump reqUires ;J V,rcn
O. (Consider the effect of
the perturbation on the equilibrium constant.)
4.6.
An "old chemists' tale" is that, near 300 K, the reaction rate doubles with
each 10 K increase in temperature. What is the "universal" Ea?
4.7.
"*
AX*
+ BX~ AX + BX*
Problems
105
Table 4.6
[PLPl o
[Glulo
[Cu 2 +1 o
(mM)
(mM)
(mM)
Initial rate
(11M sec-')
0.1
0.1
0.2
0.3
0.5
0.5
0.2
0.2
0.2
0.2
0.2
8.0
8.0
8.0
8.0
8.0
8.0
0.7
1.9
3.2
8.0
11.0
0.2
1.0
0.2
1.0
1.0
0.2
0.4
0.4
0.4
0.4
0.4
0.12
0.11
0.24
0.33
0.56
0.59
0.023
0.054
0.094
0.23
0.31
[M. E. Farago and T. Mathews, J. Chern. Soc. A., 609 (1969)]. Some results
are tabulated in Table 4.6 for 25C and pH 4.0.
(a) Deduce the rate law.
(b) What is the rate constant?
(c) What is the standard deviation in the rate constant?
4.10. Fe(III) exists in two forms in acid solution; these are readily interconverted
according to the equilibrium
(a) If the only species that reacts with EU(lI) (see Section 4.10.1) is FeOH~:q)
show that the apparent rate constant for the electron-transfer reaction
would be
k'
k*Kal([H+]
+ Ka)
(b) Test this possibility by plotting 11k' vs. [H+] using the data in Table 4.3
for T = 1.4C.
(c) Use a least-squares procedure to determine k* and Ka.
k 2([ML]
+ [L] + I/K
2)
(b) Use the data in Table 4.4 and the values of K, and K2 to compute
[ML].
(c) Test the hypothesis of part (a) both graphically and via a least-squares
procedure.
Chap. 4
106
4.12. (a) Use the data of Table 4.3 to compute ko and k-l for the Eu(I1)
reaction at 15.8OC and 25C.
(b) What is Ea for the two pathways?
+ Fe(III)
4.13. (a) Show that if for all data, Yi = kXj + q, the correlation coefficient
r (4A.8) is unity.
(b) When variables are uncorrelated the value of x has no influence on
that of y. Construct an argument to show that, when N is large, for such
variables,
O.
where R is 4,7-dimethyl-I,10-phenanthioline, has been studied using temperature jump [J. Halpern, R. J. Legare, and R. Lumry, J. Am. Chem. Soc.
85, 680 (1963)]. The experiments were carried out under conditions where
the initial concentrations of the reactants were equal, i.e., [FeR3'+]0 =
[lrCl 0 2-]0 = Co. The results suggested that the decay constant is proportional to Co.
(a) Assume a simple bimolecular reversible mechanism and show that
x 10 (sec)
(0C)
Cox 10 (M)
10
0.55
10
7.5
5
33
55
66
18
0.40
10
7.5
5
32
42
59
30
0.28
10
7.5
5
23
31
44
Problems
107
(c) The data were obtained (r "" x- 1 ) as shown in Table 4.7. Determine
k+ and k_ at the three temperatures.
(d) Determine A and Ea for both forward and reverse steps.
4.15. (a) For the system of Problem 4.14 write the second-order rate law.
(b) In an ordinary kinetic experiment with initially equal reactant concentrations Co, show that the rate of product formation is
X=
k+Co' - 2k+CoX
+ (k+
- k_)X'
K1/'
K 1/'
K 1 /'
K 1 /'
+K
= e"t
with y "" X/Co and x given by one of the formulas in problem 4.14b. This
expression can be rewritten to yield an equation for y,
yet) = ( 1
K 1 /' coth
xt
)_1
+ Cr.0 =
+ HCrO.- =
7 '-
0.018 M
2HCrO.-,
Kl
H.CrO.,
K. = 5 M- 1
(a) Show that, after equilibration of Cr(VI) and H+, but before reaction
with 1-, the hydrogen-ion concentration, h, is
150)
After R. G. Wilkins, The Study of Kinetics and Mechanism of Transition Metal Compounds (Boston: Allyn and Bacon, 1974), p. 56.
Chap. 4
lOS
Table 4.8
[H+]o
[Cr.O,~]o
[1-]
-(d[1l/dt)o
(mM)
(mM)
(mM)
(ftM sec-I)
20.0
20.0
20.0
20.0
20.0
20.0
20.0
20.0
19.8
40.0
60.0
80.0
139.7
10.0
20.0
30.0
40.0
30.0
30.0
30.0
30.0
30.0
30.0
30.0
30.0
30.0
0.14
0.55
1.27
2.13
0.052
0.76
4.97
7.10
0.45
0.80
1.02
1.20
1.65
20.0
20.0
20.0
20.0
4
16
40
49.5
12.0
12.0
12.0
12.0
12.0
(b) Set [HCrO. -]0 = x and show that [HCrO. -]0 is found by solving the
equation
2x'
+ K,(1 + K.[H+].)x
_ 2K,[Cr(V1)].
K,[H+]0(K2 x)'
1 + K 2x
(c) Since the right-hand side of the equation in (b) can be neglected show
that
[HCrO.]o- = x = 4K, [{ (1
+ K.[H+]o)' +
16[Cr(VI)]0
}'/' - 1 - K.[H+]o ]
K,
(d) Use the results from (c) and (a) and show that the rate law is d[I-]/dt
= -k[HCrO.-][I-]'[H+p.
(e) What is the rate constant, k?
4.17. The stoichiometry for the reaction in Problem 4.16 is
2HCrO.-
+ 61- + 14H+ =
2Cr 3 + + 31.
+ SH.O
and a more detailed study of the reaction kinetics leads to the rate law
(a) What experimental approach can you devise for confirming this result?
(b) Could it be established by monitoring the initial rate of the back reaction?
Problems
109
was studied using excess Fe3+ and H+ [J. H. Espenson and S. R. Helzer,
Inorg. Chem. 8,1051 (1969)). Small changes in [FeCrO.+] were monitored
with the result that
dln[FeCrO.+]
---d'--t---
k,
90.6
57.1
37.7
2.34
1.91
1.46
T (K):
500
447
404
220
210
200
+ 2Fe'+ =
Hg,H
+ 2Fe +
3
for which the forward rate law is R j = k j [Hg2+][Fe'+]. At equilibrium forward and backward rates are equal. Formulate the equilibrium constant,
determine some different expressions for the quantity [Hg'+HFe'+], and
show that Rr may have the form (among others)
Rr ex [Fe 3 +HHg, '+ ]I/2
ex [Fe3+]'[Hg2H]/[Fe H HHg H ]
ex [Fe3+]" [Hg, H ]3/2/ [Fe H ]' [Hg'+]'
(i)
(ii)
(iii)
After R. G. Wilkins, The Study of Kinetics and Mechanism of Transition Metal Compounds (Boston: Allyn and Bacon, 1974), p. 57.
Chap. 4
110
Table 4.9
Frequency
Vi
1000
2000
Vs
1500
500
Table 4.10
t (sec):
0.02
0.04
0.08
0.12
0.16
0.20
0.24
0.28
.1A(v, ) :
7.5
8.9
9.5
9.5
8.7
7.8
6.3
5.3
4.3
.1A(v.) :
13.9
12.3
10.8
8.5
6.8
5.8
4.3
3.5
2.8
given in Table 4.9, and the measured differential absorbances are given
in Table 4.10.
(a) Determine [A(t)] - [A( 00 )], [B(t)] - [B( 00 )], the path length of the cell is
0.1 m.
(b) Are the data consistent with a single relaxation process?
4.23. The rate of halogen hydrolysis X. + H 20 ~ X- + H+ + XOH, where
X = CI, Br, or J, could not be determined until the advent of fast reaction
techniques, especially temperature-jump [M. Eigen and K. Kustin, J.
Arner. Chern. Soc. 84, 1355 (1962)]
(a) Derive the expression relating the reciprocal relaxation time (l(T) to
the forward and reverse rate constants and concentrations in the halogen
hydrolysis reaction.
(b) Calculate (within 10% accuracy) the relaxation time, T, for bromine
hydrolysis when total bromine concentration is 3 x 10- 3 M; at 293 K and
ionic strength 1.0 M, Keq = [BrOH][Br-][H+]f[Br.] = 6.9 x 10-9 M2 and
k+ = 110 sec-i.
General References
Microscopic Reversibility and the Principle of Detailed Balance
R. K. Boyd, Chem. Rev. 77, 93 (1977).
S. R. deGroot and P. Mazur, Non-Equilibrium Thermodynamics (Amsterdam: NorthHolland, 1962), pp. 92-100.
D. ter Haar, Elements of Statistical Mechanics (New York: Rinehart, 1954), pp. 381-382.
A. Messiah, Quantum Mechanics, translation by J. Potter (Amsterdam: North-Holland,
lQIlIl) nn.664-675.
General References
111
Relaxation Theory
C. F. Bernasconi, Relaxation Kinetics (New York: Academic Press, 1976), Chapters 1-5.
s.
C2H 4
+ H2
----"0
C2H 6
1 + 1 + He
Br + Br + A
(1)
(2)
----"0
----"0
12 + He
Br 2 + A
113
Chap. 5
114
+ CH 2CO
A. A. Frost and R. G. Pearson, Kinetics and Mechanism, 2nd ed. (New York: John Wiley, 1961), pp. 110-111.
Sec. 5.2
115
+ k_[B]
d~]
-k+[A]
d~]
+k+[A] - (L
+ k+')[B] + L'[C]
(5.2)
L ai(O) exp(-xit),
etc.
(5.3)
i~l
where the Xi are decay constants determined by the rate constants and the
coefficients ai(O) are fixed by both the rate constants and the initial conditions.
Deduction of the mechanism (5.1) and of the rate equations (5.2)
is greatly simplified if the various decay processes occur at vastly different
rates. Analysis of the case n = 2 exhibits all chemically significant features
of the general mechanism. To introduce the ideas of time-scale separation
and of a trace intermediate assume that the reactions go to completion,
i.e., k_ - k_' == O. Then, if initially only A is present, the concentrations
of A, B, and C are found by direct integration of the rate equations
[A]
Ao exp( -k+t)
[B]
[C]
(5.4)
Ao exp( -k/t),
[C]
Ao - [B],
Chap. 5
116
+-
o....
C
Q)
u
c
80.5
Q)
>
+-
Q)
0::
2.0
1.0
3.0
time
Fig. 5.1. Relative concentrations of A, B, and C for the mechanism (5.1) assuming
k/ = 2k+ and k_' = L = O. The time scale is arbitrary.
so that [B] is always a fixed (small) fraction of [A]. Here it may be difficult
to establish that the reaction proceeds in two steps since B is present in
very low concentration at all times. Monitoring of neither A nor C would
suggest that the reaction proceeds via an intermediate; other evidence
would be needed.
If k+ ~ k/ all three species are present simultaneously as shown
graphically in Fig. 5.1. Experimentally one infers the two-step mechanism
by observing C. Were only a single first-order process important, such as
in the two limits considered, then In [Ao - C] would be a linear function
of t with a slope determined by the significant reaction. For the intermediate
case, as shown in Fig. 5.2, this is not true until t is very large.
In general, reactions are reversible. If the rate equations (5.2) are
analyzed using the matrix method of Appendix A for the special case
n = 2, the two decay constants of (5.3) are /{:
(5.6)
where
(5.7)
Sec. 5.2
117
3.0
2.0
-In(i-
[el )
Ao
1.0
OL-~--~~--
1.0
__~=-____~~__
2.0
time
3.0
Fig. 5.2. Plot of In (1 - [C]/Ao) for the mechanism (5.1) assuming k+' = 2k+ and
L' = k_ = O. The plot is nonlinear at short times since both reactions are significant.
[8]
Ao
OL--~-~~~-~~--~--~~--~
. 0.1
0.2
2.0
4.0
6.0
8.0
time - Fig. 5.3. Plot of [BJ/Ao for the mechanism (5.1) assuming k+ = 20, L = 10, k+' = 0.1,
and L' = 0.2. Note the clean separation of time scales. There is rapid growth of the
intermediate B governed by the fast decay (u+ = 30.034) which is followed by a slow decay
(u_ = 0.266) to equilibrium. The individual decay constants can be determined by analyzing the early growth and the late decay separately. Note that for 0.15 :'S t :'S 0.5, [BJ essentially is constant. At equilibrium [BJ/Ao = 0.5.
118
Chap. 5
Sec. 5.2
119
1.01------~
0.8
[A]
0.6
AO
0.4
0.2
/--~
/
0.006
y_--- -----
0.004
Ao
\[8J
[8]
AO
0.002
O~-~--~I---~--~--~-~
0.1
0.2
12
16
timeFig. 5.4. Plot of [Al/A. and [BI/A. for the mechanism (5.1) assuming k+ = 0.1, L = to,
= 20, and k_' = 0.2. These are conditions under which the steady-state assumption is
valid. There is a rapid induction period governed by the fast decay (x+ = 30.0) during
which a steady state is produced. This is followed by a slow approach to equilibrium
(x_ = 0.133). During the induction period, [A] is constant. In the steady-state domain
(t;<; 0.2), [B] is not constant; however, there is never much B present.
k+'
(5.12)
120
Chap. 5
(S.13)
Sec. 5.3
121
(5.14)
d[Oa]
dt
-k K [Oa]2
2
[0 2]
+k
-2
[0]2
2
(5.16)
which when substituted in the rate expression for either step yields the
approximate rate law
(5.17)
This, but for the factor \12, is equivalent to (5.16) if kz[Oz] ~ k_1[Oz].
The discrepancy arises because (5.15) is an approximation; the last two
terms in d[031 dtJare not exactly equal. Stated in chemical terms, the 0 +
0 3 reaction must be slow, precisely the condition for rapid preequilibraC,)
We shall see in the next section that, strictly speaking, the initial decomposition of
ozone must also involve a bimolecular step.
122
Chap. 5
tion of 03, O 2 , and O. In this example both the steady-state and preequilibrium hypotheses are consistent with the empirical rate law (5.13),
and therefore adequate to demonstrate that the mechanism (5.14) is
plausible. Unambiguous demonstration of (5.14) is more difficult since, in
the preparation of 0 3 , atomic 0 is always present.(5)
If the proposed mechanism involves many steps or if some of the
elementary reactions are third (or higher) order, exact analysis can become
quite involved. The attempt at chemical simplification proceeds similarly.
On chemical grounds one postulates that certain steps involve rapid equilibration and/or that some intermediates are in.a steady state. The predicted
rate law is then compared with the one determined experimentally. If they
agree the mechanism has passed its first test.
It is not always possible to solve kinetic problems using simple approaches. Even if the mechanism only involves a few steps the data may
not readily suggest the coupled rate laws. Chemical intuition is required
to propose a plausible mechanism which then must be tested. The natural
procedure is to integrate the mechanistic rate equations and to compare
the predicted concentrations with the observed ones. If a set of kinetic
parameters can be found that satisfactorily accounts for the data, the
mechanism is a possible one. Such calculations always involve extensive
numerical work. Thus reasonable starting points are needed, even if modern
high-speed computers are used.
While closed-form solutions for irreversible two-step mechanisms
involving combination of uni- and bimolecular steps have been formulated, (6) in general the rate laws must be integrated numerically. Straightforward approaches such as the Runge-Kutta methods are usually adequate
to insure rapid convergence. (7) The process must be repeated for each set
of rate constants to be tested. The most efficient procedure is usually to
program a computer to carry out the calculations. In some instances the
simple numerical integration algorithms converge slowly, or not at all.
This is a particular problem for the complex mechanisms that describe
systems for which there are multiple steady states, oscillations, or ex(6)
(6)
(7)
Sec. 5.3
123
<
ZAB =
Non dlBcAB[A][B]
With reasonable values for molecular parameters this yields k 2 ;:S 1010_
1011 M-1 sec- 1 when T ,....." 500 K. A similar bound can be found for k3 if,
for the reaction to occur, it is assumed that two molecules must be within
a collision diameter when struck by a third. The result is k3 ;:S 108_1010
M-2 sec- 1 (Problem 5.23). In solution the rate constant is limited by the
rate of diffusion (Sections 5.7 and 9.13).
(8)
(9)
(10)
glewood Cliffs, N. J.: Prentice-Hall, 1971), Chapter 11. For a general introduction
see D. Edelson, J. Chern. Ed. 52, 642 (1975).
C. F. Bernasconi, Relaxation Kinetics (New York: Academic Press, 1976), Chapters 3,
4, and 7; G. W. Castellan, Ber. Bunsen Gesell. 67, 898 (1963).
Chap. 5
124
cyclopropane
+O
propene
A * ~ products
An A molecule is activated by collision with another A molecule. In the
collision process the relative kinetic energy of the two colliding molecules
is transformed, in part, to internal energy of one of the molecules, thus
"activating" it. Such inelastic collisions (in which kinetic energy is not
conserved) provide the mechanical basis for understanding the activation
process. (12) The excited molecule A * is stabilized in two ways. Either it is
deexcited by another inelastic collision with A or it decomposes to form
products. The rate expressions deduced from (5.18) are
d[~*]
(5.19)
A * since molecular collisions are constantly occurring. The rates of production and depletion of A * greatly exceed its net rate of change. Hence
(11)
(12)
Sec. 5.4
Unimolecular Decomposition-Gases
125
d[product]
dt
k+kd[A]2
kd + L[A]
(5.20)
d[product]
dt
k [A]
ex
o
00
0>0
c5b
0 000
~o
cf
-1.0
of
log p (atm)
Fig. 5.5. Dependence of the isomerization rate of cyclopropane on the pressure of cyclopropane. [Adapted from H. O. Pritchard, R. G. Sowden and A. F. Trotman-Dickenson,
Proc. Roy. Soc. A217, 563 (1953).]
Chap. 5
126
A+M~A*+M
k_'
k+[A]2 + k+'[AJ[M]
L[A] + L'[M] + kd
(S.22)
which reduces to first-order kinetics when either the buffer or the reactant
is at high pressure. At intermediate and low pressures this rate law is only
qualitatively correct. Nevertheless a measure of the relative rates at which
different gases activate A is given by the "first-order" rate constant derived
from (S.22) at low gas pressure [the analogue to (S.21)],
N.
CO.
CH 4
H 2O
Toluene.
Mesitylene .
Trifluorobenzene
a
Relative efficiency
1.0
0.060
0.053
0.24
0.060
0.011
0.007
0.03
0.003
o.on 0.009
0.27 0.03
0.79 0.11
1.59 0.13
1.43 0.26
1.09 0.13
Sec. 5.5
Radical Recombination-Gases
127
The experimental rate law is often second order in X but the mechanism
cannot be
Another mode of stabilization is conceivable. Before the unstable X. molecule dissociates it could spontaneously undergo a vibrational transition and be stabilized. This is
a most unlikely process, even less common than three-body collisions in a gas. It may,
however, be important in molecule formation in interstellar space; the formation of
CH. + from C+ + H. appears to follow this pathway. For an introductory review see E.
Herbst and W. Klemperer, Phys. Today 29(6), 32 (1976).
Chap. 5
128
f\
/T(r)
/'('
, __ _
U(rl
kl
12 + 12
21+M~12+M
A third mechanism should also occur
+ OH ----' H 0
2
2CH 3 ----' C 2H 6
are intermediate cases. There are fewer vibrational modes in water (3) and
ethane (18) than in biphenyl (60); as a result decomposition of the energetically unstable molecule is rapid. However, in all cases the molecule is
Sec. 5.5
Radical Recombination-Gases
129
X+Y ~XY*
k_
XY*+M~XY+M
(5.24)
X+Y+M~XY+M
There is a two-step pathway in which the metastable XY* is deactivated
by collision with M and the parallel one-step pathway involving direct
stabilization via three-body collisions. (14) If a steady-state assumption if>
made on XY*, the predicted rate law is
(5.25)
thus the kinetics should vary from second to third order depending upon
the pressure of buffer gas and the magnitude of the rate constants. (15) For
small molecules the decomposition rate of XY* must be very fast; thus
k_ ~ k2[M] at reasonable pressures of M and third-order kinetics are found.
For larger molecules k2[M] > k_ at moderate partial pressures of M in
which case (5.25) simplifies to
Q
Since three-body collisions (in the gas phase) are very unlikely, k+ > k3[M]
unless the partial pressure of M is quite large (Problem 5.23) and (5.25)
reduces to a second-order rate law. As an indication of the effect of molecular
complexity, the observed recombination rate of I atoms is always third
(141
(151
X+M~XM,
k2'
XM+Y~XY+M
Ie '
Chap. 5
130
Ea (kJ mol-i)
Ea (kJ mol-i)
-1.7
-5.4
-6.3
-7.3
-7.1
Toluene
Ethyl iodide
Mesitylene
Iodine
-11
-10
-17
-19
a Data from G. Porter and J. A. Smith, Proc. Roy. Soc. (London) A261, 28 (1961).
order, that of phenyl radicals is always second order, and that of methyl
radicals varies from third to second order as the buffer gas pressure changes
from 0.0013 to 0.06 atm.(l6)
At low or moderate pressure the direct three-body mechanism is not
important and the apparent rate constant is, from (5.25), k2k+/L = k2K,
where K is the equilibrium constant for formation of vibrationally excited
XY. The apparent activation energy is therefore
(5.26)
where LlE(U is the energy of reaction for formation of XY and Ea(2) the
activation energy for the stabilization step. Since the first step is slightly
exothermic (a weak bond is formed), LlE(1) < 0; the second step is highly
exothermic and requires no (or negligible) activation so Ea(2) ~ O. Thus
we expect Ea ~ LlE(l) < 0, which is the rule in recombination reactions.
The same arguments apply if reaction proceeds via the weakly bonded
intermediate XM (or YM). To establish which pathway is followed experiments are carried out using a range of buffer gases. Since Ea is set by
the bond energy of the intermediate it is independent of buffer if XY*
is the transient and variable if XM (or YM) is the transient. The effect
of buffer gas in the iodine recombination is illustrated in Table 5.2; clearly
the intermediates are 1M. Comparable data for the recombination of H
atoms or of Cl atoms or for the addition of H to NO show no effect of
buffer on Ea suggesting that XY* is the likely intermediate. (17)
The mechanism (5.24) is, in its deactivation step, the converse of the
Lindemann mechanism (5.18). Whereas unimolecular decomposition poses
(18)
(17)
Sec. 5.6
131
(5.27)
where kif ""' 0.1 and k' <X e- 20 ,1001T for temperatures from 500 to 600 K.
The result defied interpretation for a number of years. Independently
Christiansen, Herzfeld, and Polanyi(20) considered the following reaction
sequence:
kl
+ Br2 ~
2Br + M
k2
Br + H2 ~ HBr + H
ka
H + Br2 ~ HBr + Br
k.
H + HBr ~ H2 + Br
k.
M + 2Br ~ Br2 + M
(1) M
(2)
(3)
(4)
(5)
(5.28)
Chap. 5
132
To determine the concentration of Br and H atoms we apply the steadystate hypothesis. Since
d[Br]
~ =
-k2[Br][H2]
+ k 3[H][Br 2] + k4[H][HBr]
I'=::::>
(5.30a)
(5.30b)
k2[H2][Br]/{k3[Br~]
+ k4[HBr]}
(5.31)
Substituting (5.32) and (5.33) into (5.29) then yields, after slight rearrangement,
d[HBr]
2k2(kl/k5)1/2[H2] [Br 2]1/2
(5.34 )
dt
1 + (k 4/k 3)[HBr]/[Br 2]
in exact accord with the empirical expression (5.27) if k' = 2k2(kl/k5)1/2
and k" = k4/k3.
It is possible to evaluate k2 since kl/k5 = Keq for the dissociation of
Br2. Then, considering the temperature dependence of k' and Keq , the
Sec. 5.6
133
activation energy for step (2) in the mechanism (5.28) is found to be 73.6
kJ mol-I, which is certainly reasonable. Since this step is endothermic by
68.2 kJ mol-I, a comparable activation energy is required.
The mechanism (5.28) is consistent with the empirical rate law (5.27).
The activation energy for step (2) is quite reasonable. However, it is also
important to show that other possible steps such as
(6)
+ H2
M + HBr
Br + HBr
M + 2H
M + H + Br
M
(7)
(8)
(9)
(lO)
--->.
--->.
--->.
--->.
--->.
+M
H + Br + M
H + Br2
H2 + M
HBr + M
2H
are not plausible. If these were significant reaction pathways the rate law
would not be (5.34). There are also other reasons for eliminating them.
The bond energy of both H2 and HBr is much larger than that of Br 2.
Thus the activation energy of either (6) or (8) is much greater than that
of step (1) of (5.28); the dissociation rate of either H2 or HBr is very slow
in comparison with the dissociation rate of Br 2. Bond energy arguments
HBr reaction (8) has an activation barrier of at
indicate that the Br
least 173.2 kJ mol-I; thus it is not competitive with steps (2) or (4) of
(5.28), each of which has a much lower activation energy. The chain termination steps involving atomic H, (9) or (lO), can be eliminated because
the H-atom concentration is too small to be effective (Problem 5.7). (22)
The H2
Br 2 reaction is an example of a linear chain reaction. In
each propagation step one atom of reactive intermediate is produced for
each atom that is used up. From (5.34) we see that, if the temperature is
constant, the rate of HBr production is greatest near the beginning of the
reaction and drops off as H2 and Br2 are consumed. If the system is not
thermostatted, the temperature will increase because the reaction is exothermic, and the reaction is thus accelerated. If heating occurs with sufficient rapidity, depletion of reactants will not be sufficiently fast to keep
the reaction in check and the conditions for a thermal explosion are present.
Br 2 reaction, because heating is relatively
While not occurring in the H2
slow, an explosion can occur in the photoinitiated H2 + Cl 2 reaction if the
light intensity is high enough.
(22)
A. A. Frost and R. G. Pearson, Kinetics and Mechanism, 2nd ed. (New York: John
Wiley, 1961), p. 239.
Chap. 5
134
Initiation:
Branching:
Termination:
+ '"
R + ...
---->.
---->.
---->.
+ ...
aR +
nR
(5.35)
The various rates r may depend upon the concentrations of any reactant
species other than the chain carrier. As long as the rate of formation of R
is small, steady-state kinetics are applicable and
so that
[R]
nrJ[r t - (a - l)rb]
(5.36)
The two mechanisms for explosion are discussed at greater length in Chapter 7.
Sec. 5.7
Simple Mechanisms-Solution
135
<
k+ ,
k_
BC* ~ B + C
+ kd)'
(24)
The solvent structure is often described as a "cage," a somewhat misleading term since
the "solvent cage" has no permanence. Different solvent molecules diffuse in and out
of the vicinity of the products so that the structure is being continuously broken and
reformed. However, the important feature which limits mobility is that the decomposition products cannot separate unless solvent molecules are displaced. The idea of
a solvent cage was introduced by E. Rabinowitch and W. C. Wood, Trans. Faraday Soc.
32, 1381 (1936).
Chap. 5
136
The corresponding rate law, assuming a steady state of the excited intermediate A *, is
d[A]
dt
(5.37)
+ B ~ (A ... B)
(A ... B)
(C ...
(diffusion)
(C ... D)
(reaction)
+D
(diffusion)
D)~C
(5.38)
A+B~(A ... B)
by solving the diffusion equation (2.23) under appropriate boundary conditions. This was done by Debye(25); kdiff depends upon the diffusion coefficents of the reactants and thus, by Walden's rule, upon the solvent
viscosity. If A and B are ions, kdiff is a very sensitive function of the solvent
dielectric constant and the product of the ionic charges. (26) In aqueous
solution most solvated species have similar size and comparable diffusion
coefficients. It is found that kdiff is most sensitive to changes in the charge
of the reacting species; it drops precipitously for reactions involving similarly
charged ions.
(25)
(26)
Sec. 5.7
Simple Mechanisms-Solution
137
Since kdiff provides an upper bound to the overall rate of reaction, any
kinetic scheme from which one deduces a rate constant greater than kdiff
must be excluded from further consideration. Thus such estimates are
extremely valuable. A listing of rate constants for diffusion-limited reactions
involving very different species (free radicals, solvated electrons, ions)
is given in Table 5.3. To be exactly comparable the data should all be taken
at the same ionic strength and temperature. At higher temperatures and
higher ionic strengths the importance of ionic charge is somewhat reduced.
Nonetheless the charge product effect remains striking. The H + + S042reaction (charge product, -2) is 200 times as fast as the OH- + HP2073reaction (charge product, +3). The effect of diffusion coefficient, for reactions involving species with the same charge product, is illustrated by
comparing the rate constant for reactions involving H + (D ""' 10- 8 m 2 sec-I),
OH - (D ""' 5 X 10-9 m 2 sec-I), and all other solvated species (D ""'2 X 10-9
m 2 sec-I). The comparison is most striking for the rates of the H+ + OHreaction and the H + OH reaction. The former is 20 times faster, which is
not an ionic effect (from the data presented, the difference in charge product
might account for a factor of 2).
An example where the criterion of the diffusion limit was used to
eliminate a possible mechanism was a study of carbonyl addition reactions
similar to that discussed as an application of temperature jump in Section
4.9. Reactions of the type
H
RNH2
I
+ R'CHO :;;::::: R-N-C-R'
I I
H
OH
H H
+ R'CHOH+
RNH
2
---->.
I I
I I
R-NLC-R'
H
OH
R'=N02~
(27)
UO.2+ + H 20
Cl- + . CH.C0 2-
CI- + . CH 2C0 2H
OH- + HP 20,8-
p.O,'- + H.O
285
298
283
293
298
298
298
0.1
0.1
0.1
0.1
10"
4.7 X 108
3xlO"
5.0 X 10"
~6x
~1.6x1010
1.2 X 10"
6.5 X 10"
1.4 X 10"
1.4 X 1010
4.3 X 1010
7x 10"
1.4 X 1011
4.7 X 1010
3.4 X 1010
a M. Eigen, W. Kruse, G. Maass, and L. de Maeyer, Prog. React. Kinetics 2, 285 (1964), where original references are given.
b J. K. Thomas, Trans. Faraday Soc. 61, 702 (1965).
C
M. S. Matheson and L. M. Dorfman, Trans. Faraday Soc. 61, 156 (1965).
a E. Hayon and H. A. Allen, J. Phys. Chern. 65, 2181 (1961).
M. Anbar and E. J. Hart, J. Phys. Chern. 69, 271 (1965).
f D. L. Cole, E. M. Eyring, D. T. Rampton, A. Silzars, and R. P. Jensen, J. Phys. Chern. 71, 2771 (1967).
g M. S. Matheson and L. M. Dorfman, Pulse Radiolysis (Cambridge: MIT Press, 1969), p. 104.
eaq + Fe(CN),'-
[Co(NH.),OH.]8+
Fe(CN)a8-
CO.2- + H.O
H+ + [Co(NH.),OH]2+
OH- + HCO.-
H+ + U0 2(OH)+
eaq + CICH.C0 2-
CaH.O- + H 20
298
OH- + CaH,OH
298
298
NH.+
298
298
293
298
~1.0x1011
(M-1 sec-I)
Rate constant
0.1
(M)
(K)
293
Ionic strength
Temperature
2 CHi:HOH ~ 2,3-Butanediol
+OH~H.O
H+ + NH.
H 20
~.
H+ + OH-
NH.OH
H.CO.
HSO.-
OH- + NH.+
H+ + HCO.-
H+ + 80. 2-
Reaction
Table 5.3. Rate Constants for Some Diffusion-Limited Reactions in Aqueous Solution
Reference
~.
~.
fl
"'"
g
~
00
......
w
Sec. 5.8
Complex Mechanisms-Solution
139
would have to have been 2.7 X 1012 M-1 sec- 1 to account for the observed
kinetics. This is 50 times as large as the rate constant for any reaction in
Table 5.3 with the same charge product. It is 400 times as large as the rate
constant for reactions with zero charge product which involve neither H +
nor OH-. Since solution reactions cannot have k ~ kdiff' this pathway
was excluded. The temperature-jump study of similar reactions carried out
9 years later provided direct confirmation of this deduction.
The solvent affects reactivity in other ways than just reducing mobility.
A case in point is the photodissociation of iodine which occurs when
iodine is irradiated with sufficiently high frequency light,
In the gas phase the quantum yield (the number of iodine molecules dissociated per quantum of light absorbed) is I; each photon dissociates a
molecule. (28) In solution the quantum yields drop dramatically, a result
that can be interpreted in terms of the effect of solvent structure. (29) Upon
absorbing light the molecule dissociates. However, separation is not immediate. The iodine atoms are trapped within a transient cage of solvent
molecules to which they may transfer any excess kinetic energy. They
undergo many collisions before separating. The observed quantum yield
is determined by competition between the rate of diffusion and recombination. It depends upon both the mass of the solvent molecule (which affects
the efficiency of energy transfer from iodine atoms to solvent) and the
viscosity of the solvent (which governs the rate of separation of the
atoms).
(28)
(29)
Chap. 5
140
+ Fe (III) =
Eu(II)
Eu(III)
+ Fe(II)
dt
k[Fe(III)][Eu(II)]
(5.39a)
ko
k_ I
[H +]
)(
[H+]
)
[H +] + Ka
(5.39b)
[H+][Fe(III)]
[H+] + K '
[F OH2+]
e
aq
K[Fe(III)]
[H+] + K
(5.40)
Assuming parallel binary reactions between Eu(II) and each of the ferric
ion species, the postulated rate law is
-
d[E~;II)]
k'[Eu(II)][Fe!ci]
+ k"[Eu(II)][FeOH~;:-]
(5.41)
ko,
k"K=k_I'
K=Ka
At 1.4C and unit ionic strength the parameters are ko = 3.38 X 103
M-I sec-I, k-I = 2.24 X 103 sec-I, and Ka ""' 7 X 10-4 M so that kIf ""'
3.2x 10 6 M-I sec-I. Under the experimental conditions (pH 1-3), [Fem >
(30)
Sec. 5.8
Complex Mechanisms-Solution
141
[FeOHi~] and only at the highest pH is the FeOH~~ concentration significant. Nonetheless, since the hydroxy complex is so much more reactive,
the pathway involving that species predominates at all but the lowest pH.
Note that neither k' nor k" are comparable to their diffusion limits (about
2 X 10 7 M-l sec-1 and 2 X lOB M-l sec-I, respectively). The chemical process
of electron-transfer controls the rate of reaction of either Fe~~ or FeOHi~.
+ V(III)
Reaction-Inhibition
+ Fe(nI) =
V(IV)
+ Fe(n)
k o[Fe(I1I)][V(II1)]
(5.42)
More detailed study indicated greater complexity. The reaction rate depends
upon the product concentration; it is accelerated by V(lV) and inhibited
by Fe (II). The complete rate law is(31)
_ d[Fe(lII)]
dt
{k'
+ k"
[V(IV)] }[Fe(III)][V(III)]
[Fe(II)]
(5.43)
(5.44)
Since the second term in the rate law is minus first order in Fe(n), it must
(31)
Chap. 5
142
involve more than a single step. In some step V (IV) must appear as a reactant. Furthermore Fe(II) must appear as a product in a reversible step in
order for it to act as an inhibitor. Recognizing that vanadium can exist
in all oxidation states from +2 to +5, a simple pathway that incorporates
these constraints is
ko
(5.45)
d[Fe(III)]
dt
species which exist in many oxidation states. There is always a further step
which reduces (or oxidizes) the intermediate to the stable state.
The first two methods are not catalytic since they are based on adding
energy to the system. The third involves gross alteration of the reaction
medium. Only the fourth, in which trace quantities of reagent markedly
affect the reaction rate, is catalysis.
Conventionally a catalyst is defined as a substance that alters the
speed of a chemical reaction without undergoing any chemical changes
Sec. 5.9
Homogeneous Catalysis
143
itself. Thus it should not appear in the overall stoichiometry of the reaction
nor should it combine other than transiently with either products or reactants. In practice this is a somewhat restricted use of the term; catalysts
often combine with reaction products or are slowly used up in side reactions
which occur along with the process that gives them their catalytic activity. (32)
A case where a catalyst accelerates reaction is the commercial process
for sulfuric acid. The first step is the oxidation of S02 to S03; uncatalyzed
this proceeds via the slow termolecular reaction
(5.46)
In the presence of NO a two-step pathway may be followed (the termolecular
NO-0 2 reaction has zero activation energy):
+ O 2N0 2
N0 2 + S02 ---' NO + S03
2NO
2 ---'
(5.47)
+H0 =
2CuCI + Pd(s) =
2
CH 3CHO
2CuCI
+ 2HCI + Pd(s)
+ PdCl
Chap. 5
144
one could imagine that by properly adjusting the conditions trace quantities
of Cu and Pd could catalyze a one-step oxidation in aqueous HCI
(5.48)
kl
B ~ products
slow (uncatalyzed)
+ catalyst -~ X
k2
fast (catalyzed)
+ B -~ products + catalyst
k2'
(5.49)
fast
kl[A][B]
+k
[A][catalyst]
so that, if k2 ~ kl' a small amount of catalyst greatly accelerates the reaction. In addition (5.49) predicts that the rate law is first order in catalyst,
a common (though not universal) experimental situation. A counterexample is the N 20 S catalyzed decomposition of ozone which is i order
in both N 20 S and 03. (36)
(35)
(36)
For their characterization see A. Aquilo, Adv. Organornet. Chern. 5, 321 (1967).
H. J. Schumacher and G. Sprenger, Z. Physik. Chern. A140, 281 (1929); B2, 267
(1929).
Sec. 5.10
Enzyme Catalysis
145
+ S , kl ,
k_l
ES ~ E
+P
(5.51)
and hence
d[P]
--crt"
(5.53)
(38)
Chap. 5
146
by plotting In [S] vs. [S] + Vt; the slope is - Km. Alternatively the reaction
velocity v = d[P]/dt can be plotted as a function of substrate concentration
in a variety of ways. The most common are
V-I
l/V + (Km/V)[S]-I
(Lineweaver-Burk)
V - Km(v/ [S])
(Eadie)
[S]/v
Km/V + [S]/V
. (Hanes)
EH+
+ S~X2~EH+ + p
1l
1l
(5.54)
1l
This particular mechanism takes into account the fact that enzymes can
be protonated which alters the specific rate constants for the various
elementary steps. Similar, but more complicated, schemes arise if either
substrate or product (or both) can be protonated. In addition substances
other than hydrogen ion may complex with enzyme, substrate, or product.
The reaction network would be analogous to (5.54). While a direct steadystate treatment of (5.54) is complicated, a systematic method for determining the rate law in terms of the specific rate constants has been worked
out.(40)
(39)
(40)
Sec. 5.11
Heterogeneous Catalysis
147
+ I,
KI
[E ][1]
[EI]
[S]
+ Km{l + [I]/Kr}
(5.55)
+ I,
KI' = [ES][I]/[ESI]
and thus compete with the product formation step in (5.51). The reaction
rate is now
d[P]
(5.56)
dt
[S]{l + [I]jKr'} + Km
Again the dependen~e on [S] is unaffected; however, the limiting velocity
at high substrate concentration is now a function of inhibitor concentration.
The rate laws (5.55) and (5.56) are kinetically distinguishable and permit
differentiation between these possible modes of inhibition. Another important, distinguishable process is noncompetitive inhibition (Problems
5.13 and 5.14).
Chap. 5
148
+ S:;:=:: A . S
A . S + ... ~ products
A
(equilibrium, K)
k2
(S.S7)
= K[A]/{l
+ K[AJ) =
bp/(l
+ bp)
(S.58a)
KjkT
(equilibrium, K')
+ K[A] + K'[BJ)
I. Langmuir, J. Am. Chern. Soc. 38, 2221 (1916); 40, 1361 (1918).
(S.59)
Sec. 5.11
Heterogeneous Catalysis
149
A . S ~ product
(5.60a)
+ B ~ product
(5.60b)
A . S
ko
A S + B
ko
S~product
(5.60c)
== -
d[A]
(Jt =
k 2 KSo[A]
1 + K[A] + K'[B]
k2KSO[A][B]
K'[B]
1 + K[A]
(5.61b)
(1
(5.61a)
+ K[A] + K'[B])2
(5.61c)
R= k 2K' S ~
K
[A]
(5.62)
(43)
150
Chap. 5
are similar which is not true. Elegant experiments, in which various gases
are used to etch polished metal spheres, have shown that different crystallographic surfaces are etched at very different rates. (44) In an extreme example,
where copper was exposed to oxygen, the {100} face reacted 17 times as
fast as the {311} face. (45) With the development of ultrahigh vacuum techniques (p ;S 10-12 atm) it has become possible to study the details of the
adsorption process. At such ultralow pressures, adsorption is slow enough
that reproducible surface crystallographic studies of an adsorbed monolayer
are feasible using low-energy electron diffraction (LEED).(46)
Depending upon the binding energy, adsorption is described as physical
adsorption or chemisorption. In physical adsorption binding is weak
(10--30 kJ mol-I) and the adsorbate-substrate separation is large, typical
of van der Waals interaction. In chemisorption, binding is stronger (>60
kJ mol-I) and the separations are short, typical of chemical bonds. The
surface structures formed by physically adsorbed and chemisorbed species
differ markedly.
Physical adsorption of inert gases on metals is studied at temperatures
between 10 and 78 K. At too high a temperature the adsorbed layer boils
off the surface. At too low a temperature an adsorbed gas molecule does
not migrate on the surface after striking it; the surface structure is random
and does not anneal to reflect the energetics of adsorbate-substrate interaction. In the temperature range for which surface equilibration occurs,
the structure of the surface layer is independent of the inert gas adsorbed
and of the metal surface exposed. Whether Xe is adsorbed on graphite, (47)
Pd, (48) Ir, (49) or CU(50) the surface layer develops a structure of hexagonal
symmetry, an arrangement which corresponds to the closest-packed plane
of crystalline Xe. Similar results are found in the binding of Ne or Ar to
the {100} face of Nb. (51) In each instance the structure of the surface layer
is determined by adsorbate-adsorbate interaction; the adsorbate-substrate
binding is too weak to have a significant effect. (52)
A. T. Gwathney and R. E. Cunningham, Adv. Catal. 10, 57 (1958).
R. W. Young, Jr., J. V. Cathcart, and A. T. Gwathney, Acta Met. 4, 145 (1946).
(46) For an introductory review see J. C. Buchholz and G. A. Somorjai, Ace. Chern. Res.
9, 333 (1976).
(47) J. J. Lander and J. Morrison, Surf Sci. 6, 1 (1967).
(4S) P. W. Palmberg, Surf Sci. 25, 598 (1971).
(49) A. Ignatiev, A. V. Jones, and T. N. Rhodin, Surf Sci. 30, 573 (1972).
(50) M. A. Chesters and A. Pritchard, Surf Sci. 28, 460 (1971).
(51) J. M. Dickey, H. H. Farrell, and M. Strongin, Surf Sci. 23, 448 (1970).
(52) The surface layer is oriented in a limited number of ways with respect to the underlying
metal, indicating some specificity in the adsorbate-adsorbent interaction.
(44)
(45)
Sec. 5.11
Heterogeneous Catalysis
151
Physically
Adsorbed State
2M-X
Chemisorbed
State
Figure 5.7. Schematic diagram for the variation of potential energy as a function of distance
from a metal surface for dissociative adsorption of a molecule X 2 Qp is the heat of physical
adsorption for the molecule, Qc is the heat of chemisorption, and QD is the heat of dissociation of the free molecule. The binding energy per chemisorbed atom is (QD + Qc)/2.
Note that for I Qc I sufficiently small, chemisorption no longer occurs. [T. N. Rhodin
and D. L. Adams, Treatise on Solid State Chemistry, N. B. Hannay, ed. (New York:
Plenum Press, 1976), Vol. 6A, p. 346, reprinted by permission.]
(physical adsorption)
(chemisorption)
(5.63)
The model was proposed by J. E. Lennard-Jones, Trans. Faraday Soc., 28, 28 (1932).
S. Anderson and J. B. Pendry, J. Phys. C. 5, L41 (1972); J. E. Demath, D. W. Jepsen,
and P. M. Marcus, Phys. Rev. Lett. 31, 540 (1973); 32, 1182 (1974).
152
Chap. 5
Appendix A
153
(5A.2)
(5A.3)
(SA.4)
-=
-AY
(SA.S)
The transformation T has decoupled the rate equations and the components
1&6)
See, e.g., H. Margenau and G. M. Murphy, The Mathematics of Physics and Chemistry
(Princeton: Van Nostrand, 1956), Chapter 10.
Chap. 5
154
(1 0 0)
1= 0 1 0
001
The roots of (SA.8) are just the Ai' To determine T multiply (SA.2) by T
which yields, with (SA.4),
dY
-=TKX
dt
T .K
-A T
(SA.9)
which is really a set of simultaneous linear equations in the unknown elements of the T matrix, lij' Introducing the expression for A, (SA.6), this
becomes
(SA. 10)
I limkmj = -A4ij
m
Ai
lij
Appendix A
155
This has three roots, one obviously A = 0, and the others are found by
applying the quadratic formula. Defining the quantities
Al
A2
Aa
+ 4Lk/P/2}j2
a')2 + 4Lk+'p/2}j2
a')2
(5A.l!)
and x_, being the smaller decay constant, governs the long-term behavior
of the system.
If, for some reason, the concentrations Xi(t) are also needed, they
may readily be computed. From (5A.4) we obtain
T-I . Y
==
5 Y
(5A.12)
t =
Chap. 5
156
X(t)
(5A.I5)
where
(5A.16)
a' )
"' -_ Ta {( 1 + -a
a' )2
4k_k/ ]1/2}
+ (-a+ -a---:
2a'
1 - -a-
2,-'-
(5B.1)
Since the terms under the square root are much less than 1 we can use the
approximation
(1
(5B.2)
)2 + 4k_k+' ] - -1 ( -2a'
- )2
a2
(5B.3)
The second term in this expansion is needed because there are quadratic
terms in the quantity being approximated. Consolidating (5B.3) yields
a (1 a
+ -'a
r=:::::; -
"'_ r=:::::;.!!...-
+1-
- a'
2k -k +')
a2
(1 + ~a _ 1+ ~a _ 2Lk+')
a
2
a
(5B.4)
a' _ k_k+'
a
Since a' ~ a by hypothesis, the term Lk+'/a satisfies the following inequalities
Problems
157
The approximation to (5.6) when k+' or L (or both) are large is found
in a similar way. Define
so that with (5.7) the expressions for the decay constants are
2r + r2+
x
=2q-{I +qr- [ I -
q
q2
4(k +k
+ k +'k-' q2
k +k ') ]1/2}
(5B.5)
Approximating the square root using (5B.2) yields
I _ ~
q
2(k+L
+ k/L' q2
k+L')
rq - k+k_ - k+'k_'
q
+ k+k_'
(5B.6)
The term k+L'/(k+' + L) has been dropped in the expression for if:
because of the inequality
Problems
5.1.
Assume a two-step mechanism of the form (5.1) for which k+ = 106 sec-l,
k_ = 103 sec-\ k+' = 103 sec-\ and k_' = 1 sec-l. Assume initially only
A is present.
(a) Calculate "+ and ,,_ using (5.6).
(b) Determine A(t), B(t), and e(t) by substituting expressions of the form
A(t) = a+ exp( -,,+t) + a_ exp( -,,-t) + a eq into the rate equations.
Chap. 5
158
5.2.
The reversible decomposition of phosgene COCl 2(g) into CO (g) and CI.(g)
has been interpreted according to two distinct mechanisms l ' )
Cl2 = 2CI
(equilibrium, K , )
+ CO = COCl
ka
COCl + CI ~ COCl2
k_a
(equilibrium, K 2 )
(i)
CI
(ii)
CI
+ Cl 2 =
CIa
+ CO ~ COCl 2 + CI
(equilibrium, K 4 )
CIa
k5
k_5
Derive the expression for the rate of formation of COCl, according to both
mechanisms. Are they kinetically distinguishable?
5.3.
+ O2 =
2N02
5.5.
The pyrolysis of ethane may be accounted for by the Rice-Herzfeld mechanism [F. O. Rice and K. Herzfeld, J. Amer. Chem. Soc. 56, 284 (1934)]:
(0 C 2H. ~2CH3
(ii) CH, + C,H, ~ CH,
(iii)
(iv)
(v)
C2H. ~ C2H 4 + H
H + C2H. ~ C2H.
H + C2H. ~ C 2H.
(initiation)
+ C'H')
+ H2
(propagation)
(termination)
(a) Show that d[C 2H 4]/dt ex d[H 2 ]/dt ex -d[C 2H.]/dt ex [C 2H.].
(b) How can this mechanism be experimentally distinguished from unimolecular decomposition?
(c) Use the bond energies for C-C, C=C, C-H, and H-H (342, 614,
415, 435 kJ mol-I, respectively) to estimate lower bounds for the activation
energy in each step.
(d) Use these estimates to show that
-d[C 2 H.]/dt
d[H 2 ]/dt
d[C 2H 4]/dt
(k,kak 4/k.)li2[C 2H 6]
(e) Estimate the apparent activation energy in the overall rate law.
I.')
159
Problems
5.6.
Another pyrolysis whose kinetics is described by the Rice-Herzfeld mechanism is the decomposition of acetaldehyde:
(initiation)
(i) CH 3 CHO ~ CHs + CHO
CHs + CH 3 CHO ~ CH 4 + CH,CHO }
(propagation)
(iii) CH,CHO ~ CO + CHs
(termination)
(iv) 2CH s ~ C,H o
(ii)
(a) Determine the rate laws for the formation of the stable products and the
decomposition of acetaldehyde.
(b) Use the bond energy data of Problem 5.5 and estimate the activation
energy for each of the apparent rate constants determined in (a).
5.7.
5.8.
Br ~ adsorbed atom
H
k7
adsorbed atom
and derive the analogue to (5.34). What reaction conditions would emphasize
adsorption?
5.9.
1
1 while at
4 33
is pH dependent. At low pH,
1
1
13
O
high pH, kObs = ko' + lO [H+] M- sec- for reaction with p-chlorobenzaldehyde [E. H. Cordes and W. P. Jencks, J. Amer. Chern. Soc. 84,
4319 (1962)]. The equilibrium constant for the protonation reaction
RCHO
H+ = RCHOH+
is K = 10- 7 26
(a) Assume the slow step in the reaction is
semicarbazide
k.
+ RCHOH+ ~ protonated
semicarbazone
Chap. 5
160
k.
I-~HIO
k_.
(equilibrium, K,)
+ Cl-
(equilibrium, Ka)
{k.K,[ClO-][I-] - L.Ka-'[IO-][Cl-]}f[OH-]
= 4.3
d[Cl-]/dt = x[ClO-][I-]/[OH-],
x = 60 sec-'
k[Fe(H.O).3+][bipy]3
2Co(II) + IONH.+
+4V(III) + 2H.0
k[Co complex]
(.8)
Hg.(I) + 2V(IV)
[Hg(II)][V(I1I)]'{(a[V(IV)] + b[V(III)J)-'
+(c[V(lV)] + d[Hg(II)])-'}
Problems
161
NO. suggest a mechanism to account for the observed rate law. (The
simplest one involves two binary kinetic steps in addition to the N,Os
equilibrium. )
5.13. When an inhibitor acts simultaneously competitively and uncompetitively,
this is known as noncompetitive inhibition. Derive the analog to (5.55) and
(5.56). Is this kinetically distinguishable?
5.14. The invertase-catalyzed hydrolysis of sucrose was studied by measuring the
initial rate of reaction as a function of [sucrose] [A. M. Chase, H. C. v.
Meier, and V. J. Menna, J. Cell Compo Physiol. 59, 1 (1962)]. The following
data were obtained(6.):
[Sucrose] (M):
Initial rate (M sec 1):
Initial rate (M sec- 1 )
in 2M urea:
0.0292
0.182
0.0584
0.265
0.0876
0.311
0.117
0.330
0.146
0.349
0.175
0.372
0.234
0.371
0.083
0.111
0.154
0.182
0.186
0.192
0.188
(a) Determine the apparent Michaelis constants and the saturation rate
for the two sets of experiments {plot (initial rate)-l vs. [sucrose]-l}.
(b) What is the nature of the urea inhibition? Estimate KJ and/or K/.
5.15. In general the reverse reaction
+ water = I-malate
= 4.7 X 10-6 M
KM' = 15.9 X 10- 6 M
/('d
After W. J. Moore, Physical Chemistry, 4th ed. (Englewood Cliffs, N. J.: PrenticeHall, 1972), pp. 418-419.
Chap. 5
162
[R. A. Alberty and W. H. Pierce, J. Am. Chem. Soc. 79, 1526 (1957)],
where V+ and V_ are the saturation velocities for the forward and reverse
reactions, respectively. Compute LlGobS and as many of the four rate constants as the data permit. (70)
5.17. Demonstrate (5.58a) and (5.59).
5.18. Consider the possibility that product desorption is significant in catalysis.
The simplest possible mechanism is
Derive the rate law for the disappearance of A and compare with the result
of Problem 5.15. Why do the results differ?
5.19. Consider the mechanism (5.60c) with the additional complication that
there is an inhibitor present which adsorbs on the surface according to the
reaction
In. + S ~ 2In . S
Kr
Derive the analogue to (5.61c). What is the apparent rate law when [In.]
is large?
5.20. In Problem 4.20, three reverse rate laws were shown to be consistent with
the stoichiometry and the forward rate law for the reaction
2Hg(II)
+ 2Fe(II) =
Hg. +2
+ 2Fe(III)
Deduce possible mechanisms to account for each of the reverse rate laws.
5.21. Laser T-jump has been used to study the reaction
After F. Daniels and R. A. Alberty, Physical Chemistry, 4th ed. (New York: John
Wiley, 1975), p. 355.
General References
163
5.22. Show that for the mechanism (5.1) under conditions where the steady state
applies, the apparent activation energy for the slow relaxation is significantly temperature dependent if
(a) at some T, k/ 0::::; L but E/
(b) A+' 0::::; A_ and E+' 0::::; E_.
*' E_.
5.23. (a) Show that the ratio of ternary to binary collisions in a gas is Za/ Z2
0::::; d/ A, where A is the mean free path.
(b) Assuming all three-body collisions are reactive obtain a numerical
upper bound for ka in a gas.
(c) At what pressure does termolecular reaction compete with bimolecular
reaction, assuming in both cases that reaction occurs with each collision?
To what density does this correspond?
General References
Gas-Phase Reactions
I. Amdur and G. G. Hammes, Chemical Kinetics (New York: McGraw-Hili, 1966), Chapter 3.
S. W. Benson, The Foundations of Chemical Kinetics (New York: McGraw-HilI, 1960),
pp. 252-264, 290--305, 308-312, 319-424.
J. Nicholas, Chemical Kinetics (New York: Halstead, 1976), Chapters 6 and 7.
A. A. Westen berg, Ann. Rev. Phys. Chem. 24, 77 (1973).
Solution Reactions
s.
Catalysis
I. Amdur and G. G. Hammes, Chemical Kinelics (New York: McGraw-Hili, 1966),
Chapter 7.
K. J. Laidler, Chemical Kinetics, 2nd ed. (New York: McGraw-Hili, 1965), pp. 256-286,
Chapter 9.
P. B. Weisz, Ann. Rev. Phys. Chem. 21, 175 (1970).
J. T. Wong, Kinetics of Enzyme Mechanisms (London: Academic Press, 1975).
Other Topics
Crystallization Kinetics:
M. Kahlweit, Ann. Rev. Phys. Chem. 27, 59 (1976).
6
Photochemistry
6.1. Introduction
In ordinary kinetic studies the required activation energy is introduced
via intermolecular collisions. When the activation barrier is large, reaction
must be run at high temperature or progress is imperceptible. Moreover,
thermal activation is quite undiscriminating. Energy is partitioned among
the various internal degrees of freedom of the reactants according to a
Boltzmann distribution. There is no procedure to concentrate the energy
of the reagents in the vibrational degree of freedom which most favors
reaction. Furthermore, thermal processes always proceed by low-energy
pathways. It is thus not possible, by heating, to select another mode of
reaction requiring vastly greater activation energy.
Photochemical activation is a powerful kinetic tool. A large amount
of energy is pumped into a molecule when it undergoes an electronic transition. Photoexcited molecules are far from equilibrium and their excess
energy is lost as the system tries to equilibrate. Such molecules are exceptionally reactive. Irradiation may accelerate reaction; it may permit reaction
by one (or many) new pathways.
A problem with ordinary photochemical activation is that too much
energy is absorbed. As a consequence the possibilities for reaction are
usually greatly augmented but there are often so many photochemical
pathways that this approach can be quite unselective. A new method, based
on vibrational excitation of specific reagent normal modes, gives promise
of allowing the excitation energy to be introduced in such a way that
enhances selectivity. By tuning a CO 2 laser to a specific absorption frequency
of the molecule being studied the vibrational energy might be introduced
in a way that favors a particular reactive pathway.
165
Chap. 6
166
Photochemistry
where
(6.2)
[Yd is the concentration of species i and ai(A) is the molar absorption
coefficient of that species. The decrease in intensity for a cell of path length
Lis IaCA) = Io(A) - I(A; L) and the attenuation attributable to a particular
species j is
(6.3)
CP;.
(6.4)
Since the quantum flux J;. is proportional to the intensity J;. = I(A)/(hco/A),
where Co is the speed of light, the absorption rate per unit volume, Ra(A),
for a beam of area A passing through a cell of volume V is
R (A)
a
~ _A_
V
hco
AAIa(A)
Vhc o
(6.5)
For a discussion of actinometry see J. N. Demas, in Creation and Detection of the Excited
State, W. R. Ware, ed. (New York: Dekker, 1976), Vol. 4, pp. 6-31.
Sec. 6.3
Spectroscopic Review
167
hv
Fe(II)
and M is the electric dipole moment operator. (2) The magnitude of the
transition moment is thus a function of the initial and final states.
There are selection rules which determine whether the transition
moment is nonzero. (3) In most photochemical applications three features
of the wave functions are of major importance. The first is spin multiplicity.
As long as the molecule to be photo excited is composed of lighter atoms
(atomic number ;S 36) spin conservation in a transition is a good approxsinglet or triplet
triplet transitions are intense
imation. Only singlet
enough to be accessible photochemically; excitation is between states in
L--'>.
(2)
(3)
L--'>.
For a precise determination of the relationship between a(Je) and the transition moment see W. Kauzmann, Quantum Chemistry (New York: Academic Press, 1957),
pp. 577-583, 645-646.
For an introductory treatment see I. N. Levine, Molecular Spectroscopy (New York:
John Wiley, 1975), Chapter 7.
Chap. 6
168
Photochemistry
the singlet (or triplet) manifold. (4) In molecules containing heavier atoms
spin conservation is no longer rigorous; it is less reliable as atomic number
increases, breaking down totally for Z <: 54.
In addition to spin, the transition moment is strongly influenced by
the geometric symmetry of the initial and final states. In molecules of high
symmetry, e.g., benzene, quinone, oxygen, relatively few electronic transitions are allowed, even within a spin manifold. A particularly important
restriction affects molecules with a center of symmetry. In these molecules
the electronic wave function either changes sign (is odd) or remains unaltered (is even) when the positions of the nuclei are inverted through the
center of symmetry. Only odd "-" even transitions are permitted.
While the occurrence of a photochemical excitation is fundamentally
governed by electronic considerations, the shape of the absorption band
and the nature of the primary photoproduct is determined by the vibrational
part of the wave functions. (5) According to the Franck-Condon principle
electronic transitions take place without significant changes in nuclear structure. A semiclassical justification follows from the fact that electrons are
very light compared with nuclei. The electron distribution may thus be
altered in an interval that is short compared to the time required for significant vibrational motion. (6)
In photochemical studies photoproducts are often characterized by
their emission spectra. The same qualitative considerations govern emission
and absorption; however, since emission does not require an energy input,
it may occur spontaneously. The larger the population of excited-state
molecules the greater the emission intensity. There are usually many ways
to deexcite a species A* ; each radiative process may be described in chemical
terms as
A*~A+hv
where k is a wavelength-dependent rate constant. The analog to (6.5) for
emission is the rate law
RiA) = k(A)[A*] = T-l(A)[A*}
(4)
(5)
(6)
(6.7)
If, as in the case of NO, the ground state is a doublet, excitation is within the doublet
manifold.
Sometimes electronically forbidden transitions are vibronically allowed, i.e., the coupling
of electronic and vibrational motion relaxes a symmetry constraint. Such transitions
are never intense. In no case does vibronic coupling affect spin considerations.
For an in-depth discussion of the Franck-Condon principle see G. Herzberg, Spectra
of Diatomic Molecules, 2nd ed. (Princeton, N. J.: Van Nostrand, 1950), pp. 194-203,
420-432.
Sec. 6.4
Primary Processes
169
T(A) is known as the intrinsic lifetime for emission. When the transition
dipole moment is large emission is rapid and T is short; for strongly allowed
transitions T is in the range I fLsec to I nsec. As the transition moment
decreases T becomes correspondingly larger until, for spin-forbidden transitions, intrinsic lifetimes of the order of seconds are not uncommon. Formulas have been developed that permit T to be calculated if the transition
is also observed in absorption. (7) If a transition is sharp, its intrinsic lifetime
is determined by the extinction coefficient and the line width. If transition
is from a broad band, as is common in photochemistry, it is no longer
k(A) that is significant but a rate constant found by integrating over the
whole absorption band. The extinction coefficient and band shape determine
the corresponding T. The Strickler-Berg result, valid for broad bands, is
where
Chap. 6
170
Photochemistry
T;
T2
~.
J
TJ
4
2
S0
So
Fig. 6.1. A simplified Jablonski diagram illustrating some pathways for energy release.
The ground state is a singlet, So. Excited states are either singlet, Si' or triplet, Ti
(i ~ 1). Vibrational excitation is indicated by a superscript v, i.e., So". Radiative processes
are denoted by straight lines, radiationless ones by wavy lines. Those illustrated are:
(1) absorption; (2) resonance radiation; (3) fluorescence; (4) phosphorescence; (5) vibrational relaxation; (6) intersystem crossing; (7) internal conversion followed by vibrational relaxation.
+ Ar ----' CO(v=O) + Ar
is < 1.8 X 10-8 at room temperature. (8) However, V-V energy transfer
occurs much more easily, as few as 10-1000 collisions being required in
near-resonant processes ;(9) the more complex the molecules the more
readily thermalization occurs by either pathway.
As a result, in solution where the collision frequency is "-' 1013 sec- 1 ,
(8)
(9)
Sec. 6.4
Primary Processes
171
excited molecules are vibrationally thermalized within their various electronic states before any other process occurs. In the gas phase large
molecules provide their own heat bath and a redistribution of vibrational
energy occurs after vibronic excitation. However, for small molecules at
low pressures collisions are infrequent; the initial vibrational energy distribution does not change unless the intrinsic lifetime is long.
6.4.2. Fluorescence and Phosphorescence
The simplest way to release the excess energy is for the molecule to
re-emit light. In resonance radiation the emitted and absorbed quanta
have the same frequency; the molecule returns directly to its initial state.
Except in dilute gases this is uncommon. More often there has been vibrational relaxation. (10) If both ground and excited state are within the same
spin manifold the radiative lifetime is short, 1 nsec ;:S T ;:S 1 f1-sec; the
emission process is termed fluorescence. When the transition required for
deexcitation does not conserve spin T is much larger, 1 msec ;:S T ;:S 10 sec;
the process is phosphorescence. All three radiative processes are illustrated
in Fig. 6.1.
In many photochemical problems the reactions of a photoexcited
species are studied. As we shall see, in order to characterize the kinetics
the measured lifetime in the absence of reaction, To, must be known. To
establish To one measures the quantum yield for fluorescence, ((Ii A),
defined as
(6.8)
This process is often called internal conversion. However, we shall limit use of the latter
term to describe radiationless singlet-singlet electronic deexcitation.
Chap. 6
172
Photochemistry
Excitation
wavelength (nm)
'Pf
-350
0.39 a
Intersystem crossing
<398
O~
Predissociation
-350
0.26c
Intersystem crossing
a E. M. Anderson and G. B. Kistiakowsky, J. Chern. Phys. 48, 4787 (1968); A. E. Douglas and
C. W. Mathews, J. Chern. Phys. 48, 4788 (1968).
~ R. G. W. Norrish, J. Chern. Soc. 1929, 1611.
C Quoted in W. C. Gardiner, Rates and Mechanisms a/Chemical Reaction (New York: Benjamin,
1969), p. 262.
rapid vib.
+ hVa~ A*(S1V)--~,
A*(S)
relaxation
1
ka
(6.9b)
Some quantum efficiencies are given in Table 6.1. Note that even at low
pressures molecules may be deexcited without fluorescing.
(11)
(12)
W. Kauzmann, Quantum Chemistry (New York: Academic Press, 1957), pp. 647-650.
J. P. Simons, Photochemistry and Spectroscopy (London: Wiley-Interscience, 1971),
p.174.
Sec. 6.4
Primary Processes
173
where Ii is the complete Hamiltonian of the system and 'lfJi and 'lfJt are the
wave functions of the initial and final vibronic states. (14) In polyatomic
molecules etCE) may be extremely large since there are many possible
(13)
(14)
174
Chap. 6
Photochemistry
final state
Fig. 6.2. Radiationless conversion in a molecule with a high density of vibronic states.
As (!j ~ (!i, k+ ~ k_. See text for estimates of (!j as a function of the number of atoms
in a molecule.
6.4.4. Dissociation
Figure 6.4 illustrates for a diatomic molecule, the possible excitedstate potential energy curves that permit dissociation. As mentioned
previously (the Franck-Condon principle) electronic transitions occur so
rapidly that the nuclei are, in effect, clamped during transitionY7) Quantum
mechanics requires that the transition occur between states which have
large probability amplitudes at the same internuclear separation. The
vibrational wave function of the ground state is maximum at the center of
the well; those of excited states have major maxima near the classical
(15)
(16)
(17)
Sec. 6.4
Primary Processes
175
---r---
T2
--r--+--- 1j
320 kJ mol-!
178 kJ mor!
312 kJ mol-!
so--L-----------~~~---
Fig. 6.3. Electronic energy le~'els in anthracene. The radiationless conversion rate SI ~ T2
is much larger than that for SI ~ Tl reflecting the effect of energy differences.
(b)
(c)
Fig. 6.4. Ground- and excited-state potential energy curves illustrating possible dissociation pathways. For a diatomic molecule r is the internuclear distance; in a polyatomic
molecule r is a normal mode displacement. Vibrational separations are greatly exaggerated.
The vertical arrows correspond to Franck-Condon allowed transitions. Where possible,
molecular examples are indicated. (a) Repulsive excited state-immediate dissociation.
(b) Crossing of bound and repulsive excited states-intersystem crossing leads to predissociation. (c) Metastable excited state-tunneling leads to dissociation. (d) Bound stateexcitation energy greater than dissociation limit.
Chap. 6
176
Photochemistry
+0
CHBra ~ CHBr. + Br
H,O ~ HO + H
Cyclobutanone ~ CO + Cyclopropane
H 0. ~ OH + OH
7 CH. + H. (preferred)
CH.
.
'" CHa + H
Octatetraene ~ Benzene + C.H.
NO,
NO
110)
For a discussion of the properties of the nuclear wave functions see G. Herzberg, Spectra of Diatomic Molecules, 2nd ed. (Princeton, N. J.: Van Nostrand, 1950), pp. 76-79,
93-94.
Dissociation and predissociation in diatomic molecules are discussed in detail in G.
Herzberg, Spectra of Diatomic Molecules, 2nd ed. (Princeton, N.J.: Van Nostrand,
1950), pp. 387-434.
Sec. 6.4
177
Primary Processes
6.4.5. Isomerization
The relaxation mechanisms that lead to photoinduced isomerizations
such as
cis-olefin + hv -' -' trans-olefin (numerous steps)
r:r + h,~~ @
(num"ou, "'p')
Chap. 6
178
Photochemistry
(6.16)
Modifying the arguments which led to (6.11) to account for the chemical
process yields
(6.17)
or with (6.12)
CPj(O)/cpAc)
(kj
+ k2ToC
(6.18)
precisely the Stern-Volmer form (6.15). It is now obvious why determination of To is important. Only if it is known can rate constants for reaction
be established; otherwise only relative rate constants are accessible. For
aqueous acridine To is 13.5 nsec and values for k2 may be calculated. (22)
There are two significant limitations in using competition experiments,
of which fluorescence quenching is representative, to establish rate laws
and mechanisms. The methods monitor the depletion of a single species,
in this case photoexcited acridine. As such they provide information about
the first step in a reaction and, by themselves, do not characterize the
1'1)
122)
Sec. 6.5
Secondary Processes
179
5,------.-----,,-----,------,
4
0.02
0.04
c(M)
0.06
0.08
+ A -" D + A*
(6.19)
where the D* is the molecule which absorbed the radiation. This process
is extremely useful; photosensitization permits the production of A *
molecules in cases where direct photoexcitation of A is either impossible
(a forbidden transition) or inconvenient. The requirements for radiationless
energy transfer are:
(1) the process is nearly resonaLf (small increments of internal energy
can be accommodated by changes in translational energy);
(23)
The reaction of amines with photoexcited acridine actually follows a multistep pathway, in part diffusion controlled. See A. Weller, Prog. React. Kinetics 1, 187 (1961).
Chap. 6
180
Photochemistry
(2) spin angular momentum is conserved (not necessary if the photosensitizer is a heavy atom).
to occur. (24) Diffusion no longer limits the rate constants; k is not viscosity
dependent and, in the most efficient aqueous processes, is ,-...,3 X 1011 M-1
sec-I, many times the ordinary diffusion limit of ,-..., 1010 M -1 sec- 1 (Section
5.7). A quantitative theory for k yields
k ex - I6
J l(v)(v)
1 dv
v
-4
where (v) is the extinction coefficient of the acceptor, l(v) is the normalized
fluorescence of the donor, and R is the distance between donor and acceptor.(25) The R-6 dependence has been directly verified by constructing
systems in which donor and acceptor are substituent groups within the
same molecule, but widely separated by a rigid, nonconjugated system. (26)
Values of R determined from k are in good agreement with those deduced
from measurements on molecular models.
Reactions such as
30*
+ 1A
30*
+ 3A
----'0
10
+ 3A*
----'0
10
+ 1A*
T. Forster, Disc. Faraday Soc. 27, 7 (1959); Naturwiss. 33, 166 (1946).
T. Forster, Disc. Faraday Soc. 27, 7 (1959).
s. A. Latt, H. T. Cheung, and E. R. Blout, J. Am. Chern. Soc. 87, 995 (1965).
Sec. 6.5
Secondary Processes
181
Hg(T2)
Hg(T2)
Hg(T2)
V)
CH 2CO(Sl V )
CH 2
(27)
(28)
(29)
(30)
~ CH 2CO
+ CH 2CO ---' C H. + CO
(vibrational relaxation)
Chap. 6
182
Photochemistry
---->.
---->.
6.5.3.2. Photolysis of N0 2
N0 2 + hv(320 < A < 397.9 nm)
O(To)
+ N0 2
---->.
N0 3 *
...--?
NO
---->. ---->.
NO (Do)
+ 02(V ;S 12)
+ O(To)
~N03
o + O2
0 3 + N0 2
---->.
---->.
0 3,
N0 3
N 20 5
03
+O
+H 0
+ NO
N0 2 + O 2
N0 3 + N0 2 :;;=: N 20 5
2,
2
---->.
---->.
2HN0 3
(31)
Sec. 6.5
Secondary Processes
183
hv
(33)
(34)
(35)
L. Kaplan, K. E. Wilzbach, W. G. Brown, and S. S. Yang, J. Am. Chem. Soc. 87, 675
(1965); K. E. Wilzbach, A. L. Harkness, and L. Kaplan, J. Am. Chem. Soc. 90,
1116 (1968); and L. Kaplan and K. E. Wilzbach, J. Am. Chem. Soc. 90, 3291 (1968).
D. E. Milligan and M. E. Jacox, J. Chem. Phys. 47,5146 (1967).
G. E. Ewing, W. E. Thompson, and G. C. Pimentel, J. Chem. Phys. 32, 927 (1960).
Chap. 6
184
Photochemistry
6.6. Chemiluminescence
Chemiluminescence is, in a general sense, the opposite of photodissociation. Here a molecule is formed in an excited state by means of chemical
reaction. The new species then fluoresces to deexcite. It is the process which
accounts for the characteristic colors of flames. An example is found in the
combustion of CO where atomic oxygen is produced, most likely as the
first step in the oxidation of CO. Then the following sequence of reactions occurs(36):
+ BC --" AB* + C
(36)
A. G. Gaydon, The Spectroscopy of Flames (London: Chapman and Hall, 1957), Chapter 6; R. N. Dixon, Proc. Roy. Soc. (London) A27S, 431 (1963).
Sec. 6.7
185
Exothermicity
(kJ mol-I)
Reaction
It
fr
Iv
H + CI.
HCI + CI
203
0.54
0.07
0.39
D + CI.
DCI + CI
208
0.50
0.10
0040
+ Br.
HBr
+ Br
183
0.39
i).05
0.56
F+H.
HF+H
145
0.26
0.07
0.67
F+D2
DF+D
144
0.25
0.06
0.69
CI + HI
HCI + I
142
0.16
0.13
0.71
CI + DI
DCI + I
142
0.15
0.14
0.71
Y\<
(Conrotatory)
X
(Disrotatory)
Similar stereochemical control is found in countless reactions, both unimolecular and bimolecular. The explanation is based upon three assump-
Chap. 6
186
'
.'
TT4
'
TT4
IT*
TT3
~400
TT3
0\
Photochemistry
cyclobutene
r\
t::\
r.\
kJ mol-I
TT2
TT
(J
TTl
ill
m
H
Fig. 6.7. Energy levels and nodal structure of one-electron molecular orbitals (MO's) which
control the trans-butadiene ~ cyc10butene isomerization. The electron density in the various n-MO's is perpendicular to the carbon skeleton. The electron density in the a-MO's
is coplanar with the carbon skeleton. The spacing of energy levels is approximate. The
dashed lines indicate the position of the nodes in the various orbitals.
(38)
Sec. 6.7
187
Conrotatory
Disrotatory
2~3
,rf-
-~4
,~
~4
Fig. 6.8. Effect of different concerted rotations about the C1-C, and C 3-C. bonds
on the 71:1-MO of trans-butadiene. Conrotation distorts 71:1 ~ 71:; disrotation distorts
71:1 ~ a. Orbital correlation follows a minimum energy pathway.
Conrotatory:
(:n2' :n4)
Disrotatory:
--->.
Now consider the two isomerizations. In the thermal process transbutadiene is in its ground electronic configuration :n12:n22. Conrotatory
motion yields cyclobutene in the ground electronic configuration a 2:n 2
while disrotatory motion leads to the highly excited configuration a 2 (n*)2;
the activation barrier for disrotatory motion forbids this path. The lowest
photoexcited state of trans-butadiene is n 12n 2:n 3 which, assuming conrotatory
displacements, correlates with the highly excited a:n 2a* state of cyclobutene.
In a disrotatory displacement cyclobutene is formed in its first excited
state a 2:n:n*. The latter requires little or no activation; the former requires
a great deal and is forbidden. The stereochemical discrimination found in
(6.20) is thereby accounted for. (40)
This is a direct consequence of quantum mechanics. Any analysis which suggests that
the ordering of the energy levels of one-electron orbitals of the same symmetry may be
inverted as a molecule distorts is approximate. Important interactions must have been
ignored if such a result is predicted. If these interactions are accounted for the levels
no longer cross. See W. Kauzmann, Quantum Chemistry (New York: Academic Press,
1957), pp. 536-539 for a discussion of a similar problem: the change from ionic to
covalent binding in NaCl as the Na-Cl separation is varied.
(,0) For an extended discussion see R. Hoffman and R. B. Woodward, Science 167, 825 .
(1970).
(39)
188
Chap. 6
Photochemistry
(42)
Sec. 6.8
Laser Activation
quasi - continuum
y= 4
y=3
--
y= 2
y=1
y= 0
189
--
------
(2,J-I) -
(3,J)
t. = h"'23+ 2Jr
(I,J-I) -
(2,J-I)
t. = hW ,2
(O,J) -
(I,J-I)
t. = hw01 - 2J r
providing enormous photon fluxes the laser, because of its very narrow
line widths, provides highly selective excitation. A properly tuned CO 2
laser has been used to selectively excite either IOBCl 3 or 11 BCI 3 (43) At
present only two other regions of the infrared spectrum are accessible with
high-power lasers-the CO and HF lasers which emit in the vicinity of
5 [Lm ("-'24 kJ mol-I) and 2.7 [Lm ("-'45 kJ mol-I), respectively.
In laser-induced gas-phase reactions the first step is absorption of a
vibrational quantum which, at moderate pressures, would only heat the
gas since V - T energy transfer occurs. If this process were dominant the
laser would be no more than a fancy (and expensive) Bunsen burner.
At low pressures conditions are more favorable for multiple-photon absorption. The photon flux is large enough so that more than one quantum
can be absorbed before significant V - T transfer occurs. The only trick
is to maintain resonance between the laser frequency and the various transitions to be excited. If only pure vibrational transitions were involved
anharmonicity would ensure that a laser tuned to the fundamental frequency
(VO->I) would be nonresonant for other transitions. However, since there
are also changes in rotational quantum number resonance may be reestablished. The resonance conditions are illustrated in Fig. 6.9.
An example of multiple-photon absorption is found in the laserinduced decomposition of D 3 BPF 3 irradiated with a continuous source
(43)
R. V. Ambartzumian, N. V. Chekalin, V. S. Doljikov, V. S. Letokhov, and E. A. Ryaboy, Chem. Phys. Lett. 25, 515 (1974).
Chap. 6
190
Photochemistry
Ea
123 kJ mol- I
(44)
(4')
K-R. Chien and S. H. Bauer, J. Phys. Chern. 80,1405 (1976); E. R. Lory, S. H. Bauer,
and T. Manuccia, J. Phys. Chern. 79, 545 (1975).
For more recent work, emphasizing the effectiveness of laser activation in the reaction
NO + 0 3 ~ products, see K.-W. Hui and T. A. Cool, J. Chern. Phys. 68, 1022 (1977)
and references cited therein.
Sec. 6.8
Laser Activation
191
The thermal reaction yields B2H6 as well. (46) A related feature of laser
photoreactivity is the effect of varying wavelength. In the series of reactions
+ HBr =
B(CHa)2Br + HBr =
BCH 3Br 2 + HBr =
B(CHa)a
+ CH 4
BCHaBr2 + CH 4
BBr3 + CH
B(CHa)2Br
the first and third are accelerated by tuning the laser to A = 10.30 [Lm
while the rate of the last two is increased when A = 9.62 [Lm. (47)
Inasmuch as laser photochemistry is a relatively unexplored area of
research, it abounds with ambiguities and contradictions. Two instances
suffice. The decomposition of CF 2ClCF 2Cl was studied using a low-power
("'-'25 W cm- 2) laser at high pressure (",-,0.4 atm).(48) The decomposition
rate was very dependent on the irradiation wavelength; it was 160 times
faster when A = 10.86 [Lm than when A = 9.51 [Lm. Since both wavelengths
are strongly absorbed and since, under the experimental conditions, there
are",-, 104 collisions between each absorption event, V-T transfer should
lead to rapid laser heating and loss of selectivity which was not observed.
Speculation surrounds possible v-v transfer which may further excite
one molecule up the vibrational ladder. Frequency matching may be better
at the longer wavelength. However, there is presently no clear resolution of
the apparent contradictions.
Large-scale utilization of laser techniques would seem to require high
pressures (",-,0.1 atm) and high intensity (",-,10 6 W cm- 2). However, under
such conditions one important feature of laser photochemistry would appear
to be easily lost-selective vibrational excitation. While thermalization
via V-T transfer is slow, collisions still affect atomic motions and work to
redistribute energy localized in one mode among the molecules' other
vibrational degrees of freedom. In CH 3 F, "'-' 100 collisions are required to
complete the energy redistribution within the molecule; however, energy
transfer between nearly resonant modes requires ::S 10 collisions. (49) The
consequence for high-pressure laser activation is apparent; at p "'-' 0.1 atm
efficient energy redistribution should require < 50 nsec. However, studies
146)
147)
148)
149)
192
Chap. 6
Photochemistry
of the decomposition of CCl 3F and CCIF 3 suggest otherwise. (50) In experiments using a SOO-nsec pulse the vibrational energy seems neither to be
thermalized by V- T transfer nor to be collisionally redistributed within
the molecules. Instead it seems to remain in the vibrational mode that was
excited initially. Again no clear resolution of the contradictory evidence
has been presented.
(52)
(53)
Sec. 6.9
193
AI
/ I'
/
..,/
I
I
Excitation
of IOBCI 3
Excitation
of IIBCI3
A(A)
5070
5040
5010
Chap. 6
194
Photochemistry
195
Problems
UF 6 (57)
Problems
6.1.
+ B ~ A * . B ~ products
kdlff
k2
kdlss
(a) Modify the arguments leading to (6.18) and show that f{!o/f{! =
1 + ykdlflToC, where y = k2/(k2 + k dISS ).
(b) The phosphorescence lifetime of benzophenone is -160 sec. If a compound quenches benzophenone triplets at a diffusion-controlled rate
(kdlfl - 10 '0 M-1 sec- ' ) what concentration is required to quench 95% of
the phosphorescence?
6.2.
+B=
A B,
+ Bo,
(l
(aK
+1-
R)/2K,
+ 2aK + 1)'/'
+ E'K[BD/(1 + K[BD
f{!o
(57)
+ E'K[B]
+ (1 - y)'K[B]
1 + kdIffTo[B]
1 - ya
Infrared sources being developed are discussed by J. J. Ewing, in Chemical and Biochemical Applications of Laws, C. B. Moore, ed. (New York: Dekker, 1977), Vol. 2,
Chapter 6.
Chap. 6
196
Photochemistry
6.4.
6.5.
6.6.
Given the information on triplet and singlet energies shown in Table 6.4,
which compounds in the table will sensitize or quench chrysene (a) fluorescence, (b) phosphorescence?
Table 6.4
Compound
Anthracene
Benzene
Benzophenone
Biacetyl
Chrysene
Fluorene
Naphthalene
Phenanthrene
Quinoline
Triphenylene
E(S,)
E(T,)
(kJ mol-I)
(kJ mol-I)
310
481
310
255
331
397
377
339
381
343
176
356
289
230
238
285
255
259
259
280
197
Problems
6.7.
6.8.
Describe the general shape of the vibronic spectrum observed for excitation
from the v = 0 level of the ground state for the potential functions of
Figs. 6.4a, b, and d.
6.9.
The phosphorescence decay of benzophenone in benzene solution is consistent with the rate law
d[T,J/dt
-k,[T,J - k 2 [T,l"
+ So' ~ So + T,'
0.001
0.005
0.01
0.015 0.5
1.0
to-sk, (sec-'):
1.3
1.4
1.3
1.6
2.9
5.0
0.9
0.7
0.9
0.7
1.1
1.1
(e) Suggest experimental conditions that will quench diffusion but permit
hopping.
6.10. What is the likely product when the substituted cydohexadiene
Chap. 6
198
Photochemistry
Table 6.5
CPFx 100
(kJ mol-')
EBbS
(kJ mol-')
CPFx 100
77.8
1.58
55.6
2.8
56.9
0.374
46.9
0.8
50.2
0.120
41.4
0.4
35.1
0.0060
28.9
0.009
+ hv
(1)
C*
(2)
C*~~C
+ C ~2C
(4) C* + P ~C + P
(5) p* ~P + hv
(3)
C*
(6)
p*
+ P ~2P
+ C~P + C
C* + P ~C + P*
(7)
p*
(8)
p*
(9)
~~P
General References
199
Show that
rpfipc
{I
rp0fipp = {I
(c) When [P] = 0.002 M, ipp/ippo ~ 2; estimate the range of the C-P
interaction.
(d) At infinite dilution rppo/ip0 ~ i. Estimate y and the extinction coefficient
ratio Ep/EC.
(e) Measured fluorescence efficiencies are
[P] x 10' (M):
3.4 4.5
rpp/ ipc:
4.0
5.7
8.0 9.0
ip/rpc:
3.5
3.8
4.2 4.6
5.5
12
5.7
6.6
12
5.1
Assume k7 and k8 are negligible and estimate k9; TC is 1.2 X 10-8 sec. (The
diffusion limit in benzene is ~ 1010 M -1 sec- 1 .)
General References
Photochemistry
P. G. Ashmore, F. S. Dainton, and T. M. Sugden, eds., Photochemistry and Reaction Kinetics (Cambridge: University Press, 1967).
A. Cox and T. J. Kemp, Introductory Photochemistry (London: McGraw-Hill, 1971).
J. P. Simons, Photochemistry and Spectroscopy (London: Wiley-Interscience, 1971).
Energy Transfer
L. Krause, Adv. Chem. Phys. 28, 267 (1975).
D. Secrest, Ann. Rev. Phys. Chem. 24, 379 (1973).
I. W. M. Smith, in Gas Kinetics and Energy Transfer, P. G. Ashmore and R. J. Donovan, eds.
(London: Chemical Society, 1977), pp. I-56.
Chemiluminescence
T. Carrington and J. C. Polanyi, in MTP International Review of Science, Physical Chemistry,
Series I, Vol. 9, J. C. Polanyi, ed. (London: Butterworths, 1972), pp. 135-171.
200
Chap. 6
Photochemistry
Lasers in Chemistry
M. J. Berry, Ann. Rev. Phys. Chem. 26, 259 (1975).
S. Kimel and S. Speiser, Chem. Revs. 77, 437 (1977).
J. T. Knudtson and E. M. Eyring, Ann. Rev. Phys. Chem. 25, 255 (1974).
c. B. Moore, ed., Chemical and Biochemical Applications of Lasers (New York: Academic
Press, 1974), Vol. 1, Chapters 6 and 7.
Nonstationary State
Mechanisms
7.1. Introduction
In the last two chapters chemical mechanisms were studied under
simplified conditions. An attempt was made to identify each elementary
step individually. Naturally this was not always possible. Competition,
reverse reaction, consecutive reaction, and inhibition all complicated the
analyses. However, by limiting consideration to systems for which the
steady-state assumption is appropriate, reasonably simple portraits of
reaction could be drawn.
Problems of practical interest such as those that arise in combustion,
explosion, metabolism, and manufacturing processes represent a more
formidable challenge. Here the system is far from equilibrium. Various
forms of feedback may operate so that a wide range of behavior can be
found. The most familiar form of feedback leads to explosion. An exothermic reaction evolves heat thus raising the system's temperature. At the
higher temperature the reaction velocity is accelerated, the rate of heat
evolution is also accelerated, and therefore so is the rate of temperature
increase. Unless enough heat can be lost to the surroundings, explosion is
inevitable.
A less familiar form of coupling involves chemical feedback. Imagine
two parallel competing processes. As we have seen, each retards the other
since both deplete the same pool of reactants. If, in addition, an intermediate produced in one pathway is an intermediate reactant in the other,
the two paths are no longer parallel and independent. Instead the inter201
Chap. 7
202
-aT =
at
(2
x 17 T + Q -d~) - V
~
- yeT - To)
(7.1)
The second term is the chemical contribution and the last one accounts
for coupling with the surroundings; ~ is the progress variable for chemical
reaction, Q is the enthalpy release per mole of reaction, (1) and y accounts
for the effectiveness of thermal coupling between system and surroundings. (2)
Solution of (7.1) requires knowing how the rate of reaction, d~/dt, depends
upon T. Assume a simple situation and consider a single, irreversible
(1)
(2)
Q~
is generalized to
Sec. 7.2
Thermal Explosion
rate of heat
evolution or
heat loss
203
r
I
Temperature
Fig. 7.1. Heat evolution curves (1, 2, 3) and heat loss curve (straight line) as a function
of temperature.
bimolecular reaction
A
k
+B~
products
dt
k(T)(Ao -
~)(Bo - ~)
(7.2)
For an extended treatment with examples see N. N. Semenov, Some Problems in Chemical Kinetics and Reactivity, translation by M. Boudart (Princeton, N.J.: Princeton
University Press, 1958), Vol. 2, Chapter 8.
204
Chap. 7
Sec. 7.3
Population Explosion-Autocatalysis
205
(7.4)
Here n chain carriers are formed in the initiation step for which the rate
per carrier is rr. In the branching step (or steps) a carriers are formed for
each one that is consumed; the rate is (a - l)rB[R]. The termination rate
is rT[RV 4 )
To illustrate the general considerations consider the hydrogen-oxygen
system for which the important elementary steps at low pressure appear
to be the following, (5) where S is used for surface and the values in parentheses represent iJH (kJ mol-I):
Initiation:
Branching:
H2+ S
{H +0,
0+ H2
Propagation:
Termination:
HO +
kl
~HS+H
k
~HO+O
(+64)
~HO+H
(- 4)
H2~HOH
{H I S
+ H
(7.5)
(-58)
k5
(5)
16)
17)
206
Chap. 7
<
rT, A is negative
and, after sufficient time, [R] reaches the steady-state limit (5.36). However,
if A > 0 there is positive feedback. The system is unstable, branching
dominates, and [R] grows exponentially. There is a population explosion
of chain carriers. In the H2
O 2 system, which is typical, the branching
and propagation sequence, for which the stoichiometry is 3H 2 O 2 =
2H 20
2H, is only slightly exothermic. Here 28 kJ are released per mole
of H 20 while !JHf"'-' -240 kJ mol-I. The bulk of the chemical heating
occurs as a consequence of termination. Since the population explosion
leads to more termination the system heats rapidly and ignition occurs.
From (7.8) the critical condition for explosion is A = 0 or with (7.6)
and (7.7)
O. Now, if (a - l)rB
Prexp(-EB/RT) ex P/2/pd 2
For a buffer-gas pressure of Po we find a critical pressure, PI,
(8)
(9)
PI ex
p/4
exp(En/2RT)/d,
Po
PI ex
p/2
exp(EB/RT)/p od 2,
Po Pr
(7.9)
Sec. 7.3
Population Explosion-Autocatalysis
207
D
Region of
explosion
region of
normal
reaction
C
B
lIT - -...
Fig. 7.2. Explosion limits for a typical branched-chain reaction. The region ABC is known
as the explosion peninsula. The curve ABCD is the boundary between explosive and
bounded reaction rates.
Ignition is expected when P > Pl' Reality is more complicated. The explosion limits characteristic of branching chains are illustrated in Fig. 7.2.
Increasing pressure with fixed temperature and composition leads to
ignition at Pl' The unexpected feature is the region of stability at high
pressure, P2 < P < P3' The reaction velocity at constant temperature and
composition is plotted in Fig. 7.3. For P < PI the rate is very slow, if not
imperceptible. Above P2 the rate is again very slow but here it is measurable,
increasing gradually until P3 is reached. The behavior near PI and P2 is a
significant indicator of a population explosion. A system on the verge of
thermal explosion does not have an imperceptible reaction velocity.
rate of
reaction
/Pz
\
\
pressure
"-
-~
Fig. 7.3. Rate of a branching-chain reaction as a function of pressure for fixed temperature
and composition. The dashed lines illustrate the effect of inert buffer gas on the first and
second explosion limits.
Chap. 7
208
Consider PI; its temperature dependence corresponds to (7.9). Increasing the volume-to-surface ratio (increasing d) or increasing buffer-gas
pressure is predicted to lower Pl. Precisely such behavior is found in the
Hz-O z , or other branched-chain systems as indicated in Fig. 7.3. Note
that were the explosion at PI thermal in nature a buffer gas would stabilize
the system as Cv would increase and as a result the heating rate would
drop (Fig. 7.1).
In order for the system to be stable at high pressure, a process that
consumes chain carriers must become significant, otherwise autocatalysis
would remain dominant. Chain-trapping generally involves a three-body step
R
+A +M
-->.
AR
+M
(7.10)
(7.11)
A = -rQ
+ (a -
l)rB - rT = 0
+ Po ex: exp[-(EB -
EQ)/RT]
(7.13)
Sec. 7.4
209
X+Y~2Y,
(7.14)
which does not describe any known chemical process but has a crude ecological analog. If A is grass, X are deer, Yare wolves, and Z are dead
wolves (or fertilizer) then deer eat grass and reproduce, wolves eat deer
and reproduce, and wolves die. In wilderness management it is well known
that animal populations oscillate rather than settling into a steady state.
As we shall see (7.14) can account for such behavior.
Before analyzing this mechanism consider the related problem, where
reverse reaction is possible,
k.
X+Y~2Y,
k_.
(7.15)
For simplicity assume that [A] has a constant value Ao which may be arranged if, initially, A is in great excess or if the system is open and A is
(13)
Chap. 7
210
d~]
---;Jt
d~~]
ka[Y] - L3[Z]
d[Y]
+ L2[Y]2
+ L3[Z]
(7.17)
If there are small perturbations about the equilibrium state the system,
as discussed in Section 4.9, relaxes to equilibrium (see Problem 7.4). The
displacement variables [X] - X o , etc., are linear combinations of decaying
exponential functions. Direct negative feedback, due to reverse reactions
like 2X ----'" X + A, keep [X] and [Y] under control no matter what the
initial conditions.
If reverse reaction is forbidden no true equilibrium exists. No direct
feedback controls [X]; instead it is kept in check by reaction with [Y].
From (7.14) or (7.17) the coupled rate equations, maintaining [A] = A o ,
are(14)
d[X]
---;Jt = k1AO[X] - k 2[XHY]
(7.18)
d[Y] = k 2 [X][Y] - k3[Y]
dt
There appears to be a steady state when d[X]/dt
case
d[Y]/dt
0 in which
(7.19)
which are substantially different constraints than (7.16). Defining displacement variables
[X] = $(t)
(14)
+ X o,
[Y]
rJ(t)
+ Yo
(7.20)
There will be a steady increase in Z but its growth is of no immediate interest. The system
(7.18) was first analyzed by A. J. Lotka, J. Am. Chern. Soc. 42,1595 (1920); Proc. Nat.
Acad. Sci. (U.S.A.) 6, 410 (1920).
Sec. 7.4
211
+ k2~1] =
ij = k2YO~
k2~(1]
+ Xo)
(7.21)
Yo)
If the displacements are small the quadratic terms can be dropped. Using
the matrix method of Chapter 5 and assuming that ~ and 1] are linear
combinations of decaying exponential functions e-J.t, the decay constants
A, are solutions of the determinantal equation
(7.22)
from which we find
A,2
+k
2 2X OYO
d~(l
Xo
+ Xo
= -d1](1 _
1]
Yo
Yo
and integrated
(7.25)
Chap. 7
212
The general features of such curves are indicated in Fig. 7.4a. They form a
set of nonintersecting closed curves as shown in Fig. 7.4b. The singularity
is the origin in the ~-'Y) plane. Trajectories do not terminate at this point.
Instead a system displaced from the singularity oscillates in perpetuity.
Furthermore the larger the initial perturbation the larger the amplitude of
the oscillations. No relaxation occurs, even for large amplitude perturb ationsY5) What is not apparent from Fig. 7.4 is the oscillation period. However, by expanding ~(t) and 'Y)(t) in Fourier series the oscillation frequency
(7.23) may be shown to be a general feature of the rate law (7.18); w does
not depend upon the perturbation amplitude. (16)
The model mechanism predicts undamped oscillations in the concentrations of intermediates X and Y. The concentrations of reactant A and
product Z do not oscillate; however, their rates of consumption and formation do. Three conditions were needed to generate such behavior:
(i) continuous flow of matter into the system ([A] is constant);
(ii) a mechanism with coupled indirect feedback steps;
(iii) a system far from equilibrium (some steps are irreversible).
The constraints appear to be general. (17) Unlike explosion where feedback
leads to direct acceleration of reactant consumption, oscillatory mechanisms
are more subtle. According to (7.18) increasing [X] when [Y] is small
(15)
(16)
(17)
The approach outlined, considering the trajectories in the !;-'Y} plane instead of !;(t)
and rl(t) individually, has numerous applications. For an introduction to the theory of
trajectory classification see D. A. Sanchez, Ordinary Differential Equations and Stability
Theory (San Francisco: Freeman, 1968).
A. J. Lotka, J. Arn. Chern. Soc. 42,1595 (1920); Proc. Nat. Acad. Sci. (U.S.A.) 6, 410
(1920).
A discussion of the requirements for oscillatory behavior is given by G. Nicolis and J.
Portnow, Chern. Rev. 73,365 (1973); see particularly Sections L D. and lILA. A more
qualitative treatment is given by H. Degn, J. Chern. Ed. 49, 302 (1972).
Sec. 7.4
I.I
I
I
I
I
I
I
213
Xo
I
I
L __
X--+
(0)
( b)
Fig. 7.4. (a) General features of trajectories (7.25). (b) Detailed curves showing three trajectories of the concentration displacement variables ~ and 'Y} for different initial displacements.
The steady state is the origin of the ~-'Y} plane. It is inaccessible. [Figure 7.4b adapted from
A. J. Lotka, J. Am. Chem. Soc. 42, 1595 (1920).]
In a closed system reactant concentration must decrease. Thus perpetual oscillations in the concentration of intermediates are not possible.
However, if initially A is in great excess and conditions (ii) and (iii) are
met, very slowly damped oscillations may be observed. In the remainder of
this chapter we discuss some specific examples. (18) Figure 7.5 illustrates
allowed and forbidden chemical oscillation. Detailed balance forecloses
oscillations about equilibrium, as shown in Fig. 7.5a. Oscillations about
the steady state, as shown in Fig. 7.5b, violate no thermodynamic principle.
While the Lotka mechanism exhibits many features found in chemical
oscillators, it is misleading in one important qualitative respect. According
to Fig. 7.4b the amplitude of the concentration oscillations is determined
by the magnitude of the initial perturbation. Such behavior is apparently
not characteristic of real systems. Rather, independent of the initial values of
(18)
For a catalog of chemical, thermochemical, electrochemical, and biochemical oscillators see G. Nicolis and J. Portnow, Chem. Rev. 73, 365 (1973), Section II.
Chap. 7
214
(a)
(b)
Fig. 7.5. (a) Forbidden and (b) allowed oscillations in the relaxation of a concentration
displacement, 6X, as a function of time.
(20)
Two such mechanisms, presumed to describe specific oscillating systems, have been
formulated. For details see P. Glansdorff and I. Prigogine, Thermodynamic Theory
of Structure, Stability, and Fluctuations (New York: Wiley-Interscience, 1971), Chapter
14; I. Prigogine and G. Nicolis, Quart. Rev. Biophys. 4,107 (1971); R. J. Field and R.
M. Noyes, J. Chem. Phys. 60, 1877 (1974).
B. P. Belousov, Sb. Ref Radiats. Med. (Medgiz, Moscow) 1958,145 (1959).
Sec. 7.5
Belousov-Zhabotinskii Reaction
215
Malonic acid
Concentration (M)
0.013-0.5
KBrOa
0.013---{).063
Ce(NH.MNO a).
0.000 1---{).0 1
H.SO.
KBr
0.5-2.5
Trace (;S 2 x to-)
a R. F. Field, E. Karas, and R. M. Noyes, J. Am. Chern. Soc. 94, 8649 (1972).
when malonic acid was replaced by other organic acids. (21) Another example
of oscillation is the 103 -/1 2 catalyzed decomposition of H 20 2 The production of O 2 is not continuous but proceeds in bursts that are correlated with
oscillations in [1 2 ].(22) Since oxygen is produced, considerable care was
required to demonstrate that the oscillations reflect homogeneous processes. (23)
Both the Belousov-Zhabotinskii and the Bray-Liebhafsky systems have
been characterized mechanistically. (24) We consider the Belousov reaction
in detail. Oscillatory behavior is not limited to a narrow range of concentrations. Initial conditions which lead to periodic concentration variation are
given in Table 7.1. The period of oscillation is quite dependent upon initial
concentrations, as suggested by the Lotka result (7.23); it varies from ,....." 15
to ,.....,,200 sec in different experiments. Not only are there oscillations in
the [Ce(IV)]/ [Ce(III)] ratio but [Br-], which is present in trace quantities,
oscillates as well. Typical results are shown in Fig. 7.6; the amplitudes
diminish only very gradually over dozens of cycles. For the experiment il('1) A. M. Zhabotinskii, Dokl. Akad. Nauk. SSSR 157, 392 (1964); Bio/izika 9,306 (1964).
(") W. C. Bray, J. Am. Chern. Soc. 43, 1262 (1921).
(.a) W. C. Bray and H. T. Liebhafsky, J. Am. Chern. Soc. 53, 38 (1931); H. T. Liebhafsky,
J. Am. Chern. Soc. 53, 896, 2074 (1931); H. T. Liebhafsky and L. S. Wu, J. Am. Chern.
Soc. 96, 7180 (1974). Liebhafsky studied the reaction at two widely separated periods
in his career; the papers cited represent only a fraction of his work.
('.) Belousov reaction: R. J. Field, E. Karas, and R. M. Noyes, J. Am. Chern. Soc. 94,
8649 (1972); Bray-Liebhafsky reaction: K. R. Sharma and R. M. Noyes, J. Am. Chern.
Soc. 98, 4345 (1976).
Chap. 7
216
1000
500
[Ce(lV)]
~ 200 A 1 0 9 - - i 100
[Ce(III)]
50
20
, 10 B
e
L
to
~I~U
2
1 C
Ql
Ql
Cl
SECONDS .....
Fig. 7.6. Periodic behavior of [Br-] and [Ce(IV)]/[Ce(III)] in the Belousov reaction [R. J.
Field, E. Koros, and R. M. Noyes, J. Am. Chern. Soc. 94, 8649 (1972), reprinted by permission of the American Chemical Society].
3BrCH(COOH)z + 4CO z
+ 2HCOOH + 5H zO
(7.26)
(A)
Br0 3 - + HMa
Ce(IV)jCe(IlI) catalyzed
'---------.....>.> HCOOH + CO z + BrBr- inhibited
(7.27)
(B)
Sec. 7.5
Belousov-Zhabotinskii Reaction
217
Sequence (A) requires Br- as a reactant; Br- also acts indirectly to inhibit
sequence (B). When Br- is depleted reaction switches from (A) to (B).
However, (B) replenishes Br-, becoming self-inhibitory and reinitiating
(A). Oscillations arise because of the switching.
The individual steps are complex and will only be outlined. (25) Reduction of Br03 - to Br2 proceeds by consecutive removal of oxygen (twoelectron transfer)
(a) Br(b)
(c)
3-
(7.28)
(7.29)
is sequence (A)
(7.30)
The slow step in the process is (a). None of the oxybromine intermediates
are thermodynamically plausible reactants in one-electron steps such as
are required to oxidize Ce(III).
A way to reduce Br03 - in one-electron transfers postulates the Br0 2
radical as intermediate,
(7.31 )
+ 2(f) is autocatalytic
+ 3H+ + HBr0 2Ce(lV) + H 20 + 2HBr0
+ Br0
3-
2 -----'
(7.32)
It might appear that (C) would rapidly deplete Br03 -, however, as HBr0 2
disproportionates
(7.33)
(25)
The analysis is due to R. J. Field, E. Karas, and R. M. Noyes, J. Am. Chem. Soc. 94,
8649 (1972), where corroboratory details are given.
218
Chap. 7
The individual steps which regenerate Br- and Ce(lII) and complete
the cycle have not been as precisely characterized. However, Ce(IV) can
oxidize HMa or BrMa in a series of rapid one-electron transfers; the net
result is
(E) 6Ce(IV)
+ CH 2 (COOH)2 + 2H 20
(F) 4Ce(IV)
+ BrCH(COOH)2 + 2H 20
---' 6Ce(III)
+ HCOOH
+ 2C0 2 + 6H+
(7.35)
The set of processes (D), (E), and (F) comprise sequence (B).
Now consider a complete cycle in the oscillatory regime as illustrated
schematically in Fig. 7.7. At a, [Br-] is large; the major pathway is (A)
so [Br-] decreases. The ratio [Ce(lV)]j[Ce(III)] drops as reactions (E) and
(F) continue. At (3, [Br-] has become too small to sustain sequence (A);
reaction switches to sequence (B) and [Br-] plummets due to reaction
(b) reaching point y. Then the [Ce(lV)]j[Ce(III)] ratio grows due to (C)
and, as more Ce(IV) forms it replenishes [Br-] via (F). At <5, [Br-] is large
enough to allow switching back to (A); sequence (B) stops abruptly; the
remaining Ce(lV) continues to oxidize HMa via (F) and [Br-] rises sharply.
The system has returned to a.
Our analysis of the Belousov reaction has just considered the oscillatory
domain. Whether sequence (A) or sequence (B) is followed, initially there
must be an induction period before oscillation commences since periodic
behavior is triggered by replenishing [Br-], which cannot occur until
[BrMa] is large enough so that (F) becomes important. The behavior
illustrated in Fig. 7.6 is consistent with this interpretation. After initiation
Sec. 7.5
Belousov-Zhabotinskii Reaction
I [CeCIV)]
09 [CemnJ
219
(a)
(a)
(ct)
(Ii,a)
~
(tJ. y)
time
,.
[Be] drops rapidly and remains small for a considerable time until it
rises abruptly signaling the onset of oscillations.
The Belousov reaction satisfies the prerequisite conditions for oscillation. The consumption of reactants is slow; there are irreversible steps so
the system is far from equilibrium; there is delayed feedback, in this case
both catalytic and inhibitory. The concentrations of intermediates oscillate
around the steady state, which is a singular point that is never attained.
There are no oscillations in the concentrations of reactants or products,
only in their rates of consumption and formation. Evolution seems Characteristic of a limit cycle; the initial concentrations of the intermediates appear
not to affect the limiting amplitudes of the stable concentration oscillations. (26) The effect. of perturbing the oscillations by adding extra Bror Ce(IV) has been studied; after a short lag time the oscillations revert
to their preperturbation behavior, again indicative oflimit cycle behavior.(27)
A simplified mechanistic scheme, based upon the reactions outlined in this
section, has been analyzed using numerical methods such as those outlined
in Section 5.3. The time dependence of the various concentrations calculated from the model is in good agreement with experiment; the numerical
solutions appear consistent with limit cycle behavior.(28)
(26)
(27)
(28)
R. J. Field, E. Koros, and R. M. Noyes, J. Am. Chem. Soc. 94, 8649 (1972).
V. A. Vavilin, A. M. Zhabotinskii, and A. N. Zaikin, in Biological and Biochemical
Oscillators, B. Chance, E. K. Pye, A. K. Ghosh, and B. Hess, eds. (New York: Academic Press, 1973), pp. 71-79.
R. J. Field and R. M. Noyes, J. Chem. Phys. 60, 1877 (1974).
220
Chap. 7
(29)
(30)
(31)
A. M. Zhabotinskii, Dokl. Akad. Nauk SSSR 157, 392 (1964); Bio/izika 9, 306,
(1964).
A. T. Winfree, Science 175, 634 (1972); Sci. Amer. 230(6), 82 (1974). The latter article
has particularly pretty photographs.
A. M. Zhabotinskii, in Biological and Biochemical Oscillators, B. Chance, E. K. Pye,
A. K. Ghosh, and B. Hess eds. (New York: Academic Press, 1973), pp. 89-95.
Sec. 7.6
Ce(lV)
C'OllI0
(a)
221
Br-
(b)
Fig. 7.8. Schematic diagram showing development of concentration waves in the Belousov
reaction. (a) Sequence A is triggered in central domain; diffusion flows are indicated.
(b) At a later time sequence B dominates in central region; rings of concentration variations propagate outward.
222
Chap. 7
the colloidal particle the more charge it may carry and the more counterions
it will attract thus accelerating its growth and ultimate precipitation. Then,
due to diffusive coupling, the characteristic rings are formed. The significant
feature of this model is that spatial inhomogeneities form before precipitation. as seems to occur. The supersaturation explanation requires both
to be simultaneous.
(7.36)
with iron being the negative electrode. Iron is not required; similar results
are obtained using aluminum or magnesium. The oxidizing agent may also
be discarded although the "heartbeat" is then less powerful.
When there is Fe-Hg contact the iron is oxidized, Fe ---->. Fe2+ + 2e-,
and solvated Fe2+ diffuses into solution. Electrons are transferred to the
Hg surface where a cathode reaction occurs; as a result negative charge,
due to ions adsorbed on the surface, builds up. As long as there is contact
between the iron and the mercury the surface charge continues to increase;
it is apparently only limited by the rate of ion desorption from the surface.
However, the surface charge, being negative, is repelled by the negatively
charged iron electrode. If the wire does not penetrate the liquid too deeply
the Hg drop can deform sufficiently for electrical contact to break. The
adsorbed negative ions desorb into solution; the liquid resumes its spherical
shape and the circuit is reestablished. Continuous repetition leads to
pulsation.
(35)
This curious phenomenon was discovered by G. Lippman, Ann. Phys. 28 149, 546
(1873).
(36)
S.-W. Lin, J. Keizer, P. A. Rock, and H. Stenschke, Proc. Nat. Acad. Sci. (U.S.A.) 71,
4477 (1974).
Sec. 7.6
223
+ O 2 ----'- R + H02
ROO
{ R + O
ROO + RH ----'- ROOH + R
----'- RO + HO
{ ROOH
HO + RH ----'- R + H 20
RH
----'-
(7.37)
R +R
H0 2 is unreactive (see Section 7.3) and the chain carriers are R, ROO,
RO, and HO. The effect of buffer gas on energy transfer is kinetically
unimportant since the experimental conditions conform to the high-pressure
regions for unimolecular decomposition and bimolecular recombination.
As temperature is increased an unusual effect is observed; the reaction
rate drops. Thus explosion can be quenched by raising the temperature.
A new pathway that reduces the chain-carrier concentration must become
significant. A plausible candidate is the decomposition of the weakly bound
peroxy radical
ROO ----'- R + O 2
followed by olefin formation, e.g.,
(7.38)
This sequence both reduces branching by forming less ROO and increases
termination by forming more H0 2 At still higher temperatures H0 2
itself becomes reactive and may lead to propagation and branching. Thus
the reaction rate will again rise.
In the temperature domain where reaction rate falls as temperature
increases, thermal oscillations are observed for systems at pressures below
the explosion limit. The temperature of the burning gas is a periodic function of time. (38) The temperature may increase by as much as 200 K in a
(371
(381
Chap. 7
224
pulse. While it might seem that the phenomenon is an indication of inhomogeneity in the reactor, in fact oscillation is more regular and less
damped in a well-stirred system. In closed reactors only a few pulses are
observed before depletion damps the oscillations; in a continuous-flow
reactor pulsation continues indefinitely. (39)
From (7.1) the temperature-time profile in a homogeneous reactor is
dT
dt
QV
Cv
.!!:I
_ Y (T- T.)
dt
0
(7.39)
Sec. 7.6
225
As well as being of biochemical interest glycolysis is commercially important; it is used in the production of bread, beer, and wine. For a CO'lstant
rate of glucose input, oscillation is observed in the rate of CO 2 production
as well as in the concentrations of many reaction intermediates. (40) There
are multiple feedback loops; the control mechanism is far more elaborate
than in any of the physicochemical oscillators discussed previously. We
shall sketch some of the results to indicate the complexity of even a rather
simple biochemical pathway. (41)
The overall reaction (7.39) involves at least 10 separate intermediates.
The various steps, requiring C-C bond cleavage or dehydrogenation,
require mediation by a minimum of six enzymes working together. Feedback
is affected by variation of pH and of the concentration of the reduced
dinucleotide, NADH. (42) Oscillations in concentration have been found for
at least 16 separate species involved in glycolysis. The oscillators fall into
two categories. In each group the maxima and minima in the concentrations
occur at the same time. However, the two groups oscillate out of phase,
an observation that has been used to provide insight as to some features
of the feedback loop. Furthermore, the character of the oscillations changes
as the rate of glucose input is varied. At very low input level no oscillations
are found. Above a critical input level doubly periodic oscillations of small
amplitude and frequency develop. As the input rate increases further the
multiple periodicities wash out. The amplitude and frequency of the stable
oscillations increase. Finally at very high input levels oscillations are again
damped.
Of these phenomena, three have simple rationalizations. At very low
input levels the biochemical system is not sufficiently displaced from
equilibrium for stable oscillations to form. Increasing the input level
increases the concentration of the fundamental reactant which, in terms
of the Lotka model, requires an increase in frequency from (7.23). Too
great an input rate finally saturates the intermediate enzymes in a fashion
similar to that outlined in Section 5.10. As a result feedback controls are
no longer effective and oscillations damp out.
(40)
(41)
(42)
There is an enormous literature on the subject. For representative work see Biological
and Biochemical Oscillators, B. Chance, E. K. Pye, A. K. Ghosh, and B. Hess, eds. (New
York: Academic Press, 1973), Part III.
For an analysis of many of the features of the metabolic pathway see A: Boiteux and B.
Hess, in Biological and Biochemical Oscillators, B. Chance, E. K. Pye, A. K. Ghosh,
and B. Hess, eds. (New York: Academic Press, 1973), pp. 243-252.
NADH is the reduced form of nicotinamide-adenine-dinucIeotide.
226
Chap. 7
(44)
Sec. 7.7
227
(7.41),
~;
2k+(T)(C -
~/2) - 2L(T)~2
(7.43)
In (7.42) and (7.43) the quantities under experimental control are 10 , To,
and C. Increasing the radiation dose has two direct consequences-radiation
heating and increased heat loss to the surroundings-and one indirect
one-heat absorption as more N02 forms.
In a uniform steady state V 2 T, aT/at, and a~/at are all zero. Since the
parameters in (7.42) and (7.43) are known the equations, which are nonlinearly coupled due to the temperature dependence of the rate constants
k+ and k_, may be solved for ~ (or T) as functions of 10 , To, and C. Typical
results, for the same value of C, are presented in Fig. 7.9. As the surroundings cool from 250 K to 230 K the structure of the solutions changes
dramatically. At higher temperatures there is a unique N02 concentration
for each value of 10 , However, as To decreases a critical point is reached
below which there are three possible values of ~(Io, To). Two represent
stable steady states; one is unstable.
Consider the 230 K curve in detail for a system initially in steady
state A. As 10 increases ~ increases until the marginal stability point B is
reached where d~/dlo --" 00. Further increase of 10 requires an abrupt jump
in ~ to point C if a steady state is to be maintained. Yet higher-power levels
again lead to a gradual increase of ~. Now consider reducing the radiation
20
Percent absorption
(proportional to
[N0 2
10
l)
Fig. 7.9. Steady-state solutions of (7.42) and (7.43) for various choices of external temperature To. Note the hysteresis in the 230 K curve. [Adapted from C. L. Creel and J. Ross,
J. Chern. Phys. 65, 3779 (1976).]
228
Chap. 7
Fig. 7.10. Comparison of ideal (solid line) and observed (dashed line) hysteresis in the
NO,-N,04 reaction.
intensity. The system traces a path DCEFA since it is only at E that marginal
stability occurs. The net result is a hysteresis loop as 10 is varied. Above
240 K there is no hysteresis; ~ varies continuously with 10 , The set of
curves for different values of To look just like isotherms of a van der Waals
gas (turned on their side) above and below a critical point. Hysteresis
below the critical temperature has its analog in supercooling and superheating. (45)
Experiment confirms the predictions of Fig. 7.9. Above the critical
value To there is no hysteresis. Below To hysteresis is observed; however,
the jump from one branch to the other is not as abrupt as predicted. (46)
A comparison of observed and predicted hysteresis loops is shown in Fig.
7.10. Transition occurs before the marginal stability point is reached.
Under the experimental conditions there is sufficient inhomogeneity to
account for this effect. The temperature of the gas is probably not uniform
throughout the sample so that observation is of a set of systems undergoing
sequential transitions between the branches.
If uniformity were established one would still expect deviations from
the ideal hysteresis loop. Fluctuations about the steady state could induce
transitions from one branch to the other before the marginal stability point
is reached. Such behavior is analogous to nucleation of a supercooled
liquid; freezing then occurs abruptly before the limit of metastability is
reached. Referring to Fig. 7.9, one might wonder whether there could be
an analog to the Maxwell construction in the van der Waals fluid. While
a definitive answer is not yet available, the tentative answer is yes. Since
(45)
(46)
229
Problems
Problems
7.1.
(a) From Fig. 7.1 it is clear that Tc is the temperature at which thermal
explosion is inevitable. Use (7.3), assume an Arrhenius form for k(T),
and show that Tc - To = RT'/Ea
(b) Compute Tc for a reaction where Ea = 300 kJ mol- l when the system
is in a thermostatted furnace at (a) 600 K, (b) 800 K, (c) 1000 K.
(c) Why does Tc - To increase as To increases?
7.2.
The seven elementary steps of (7.5) and (7.11) account for kinetics in the
H 2-0 2 system below the third explosion limit.
(a) Show that at the first limit ks = 2k2[02]/O + k./k.[H2D and thus,
as volume-to-surface ratio increases, ks = 2k2[02].
(b) At the second limit show that k 7[M] = 2k2/(1 + k./k.[H 2]) which,
for a large reaction vessel, reduces to k7 [M] = 2k 2.
7.3.
In a stoichiometric mixture of
the second explosion limit is
P2 (atm):
T(K):
H2
0.043
0.066
750
775
0.104
800
0.156
825
(47)
Chap. 7
230
(b) Assume Z = IZo], i.e., that product is always drawn off so that C = o.
Use the matrix method of Chapter 5 and show that the decay constants
are solutions of the equation
A' - [LlXo
Yo)
+ k.k.Yo
= 0
(c) Show that both roots are positive and therefore that the system relaxes
to equilibrium without oscillation. (The roots of the equation A' - BA + C
= 0 satisfy the relation AlA. = C.)
+ 1')'/Yo = const,
7.5.
7.6.
A~X,
+ Y -..:.,., 2Y,
~'/Xo
ka
Y~Z
and thus that the system exhibits damped oscillations as it approaches the
steady state if Yo < 4Xo, i.e., if 4k.' > klk.A o.
(d) Can you give a physical rationalization for the condition governing
the onset of damped oscillation?
7.7.
[Y]
= k.[XHY] -
k.[Y]
+ Dx 17'[X]
+ Dy V"[Y]
(a) Show that the linearized rate equations for the displacement functions
Problems
231
w0 2
)l"
k.
B+X~2X+Z,
X+Y~P,
0=
+ ,1.
-k 2 Y O
k1Ao - k.Xo
-k1Ao - k.Xo + A
kaBo
0
fkG
-k5 + A
(g) The critical condition for onset of oscillation occurs when ,1. = O.
Show that, assuming Ao = B o , the constraint is Xo ~ kaBoa/[k.(a - b)].
7.9.
= k+(2C -
n-
2k_~'
+ D J72~
+ ~ V', + Qx)V/Cv
(dolx/V
-k+x - 4k_~ssx
where k+'
==
yT
+ D V'x + (2C -
(dk+/dT)Tss, etc.
~ss)k+'T - U:sL'T
Chap. 7
232
2
(b) Show that k + , = E. i +lk + ( Tss )/ RTss,
etc.
(c) Define displacements of wave vector q: T
= J(t)
cos q r and x
get)
(d) Show that the decay constants are A = {(a + d) [(a -d)2+4bc]'/2}/2.
(e) The parameters a,b,c, and d are always positive. However, as Q<O
(the reaction is endothermic), e < O. Are there value of q which permit
either damped or undamped oscillation about the steady state? Is it
possible that }.._ < O? What is the physical significance of a negative }..?
k
+F~
2P i
1
Ti
[Pd
(i
+ ki[Pd[F],
1,2
(~
~)
T2
T,
General
Referel1c~s
8
Single-Collision Chemistry
8.1. Introduction
Even if the rate law and the mechanism of a reaction have been determined there are still many unanswered questions of chemical interest.
The rate law only accounts for the overall phenomenology. The mechanism
provides a model for the basic steps in the reaction. The molecular properties which determine whether or not reaction occurs can only be inferred
from other chemical knowledge. Our previous discussion leads to only two
very general, and hardly informative, conditions for reaction: molecules
must interact (collide) and must have sufficient energy.
There are many other questions which can be posed. How is a theory
for computing a rate constant to be constructed? How may the orientational requirements for a reactive collision (such as the familiar Walden
inversion) be studied directly? Can reaction take place in a glancing collision or must molecules always collide head-on? Does reaction require the
formation of a long-lived complex or can reaction take place as molecules
fly past one another? If molecules have sufficient energy, will they invariably
react? What molecular properties promote catalysis?
These and other questions can only be investigated by isolating each
elementary step and studying the dynamics of the interacting molecules.
For gas-phase reactions, which usually proceed via binary collisions, the
chemistry of single collisions can be studied using molecular beams thus
providing new perspectives. For solution kinetics the problem is more
complicated since the basic reaction step is not the collision of two molecules
but the trapping of the reactants in a cage of solvent. Gas-phase studies
in which a molecule is surrounded by a few solvent molecules provide a
233
Chap. 8
234
Single-Collision Chemistry
way to isolate a solvated molecule from the bulk solvent. Some molecular
beam investigations of this type have been carried out.
The inherent distinction between gas and solution kinetics must be
part of any successful theory of chemical kinetics. In particular the mechanics of gas-phase reaction depend upon forces involving two, or at
most three, molecules at a time. In solution the complicating effects due to
interaction with a number of surrounding solvent molecules must also
be considered.
In this chapter we focus upon the dynamics of the reactive collision
process in the gas phase. To this end we first show how the basic measurable
parameter in collision chemistry, the reaction cross section, may be related
to its thermal counterpart, the rate constant. We then discuss the experimental approach to measurement of reaction cross section as well as a
naive model. The remainder of the chapter is devoted to the interpretation
of a number of experiments in terms of simple dynamical models for reactive
collisions. We make no attempt to develop a theory of reaction cross
sections.
+ BC----'AB + C
The A and BC molecules are constantly colliding with one another, but
not all such collisions can lead to reaction. If A collides with the C end of
BC, AB bond formation is probably much less likely than if collision is
with the B end. If the reaction is endothermic, energy must be supplied
in order for reaction to take place. The necessary energy is supplied by
collision and then redistributed among the reactants. Thus, when molecules
collide, we can expect that there is a probability of reaction (less than I)
which depends upon their relative kinetic energy, their relative orientation,
and possibly other factors which have not yet been considered.
The rate of reaction of A atoms per unit volume can be expressed as
_ dnA _
dt -
foo d
0
(8.1)
Sec. 8.2
235
where ft is the reduced mass. (2) The bracketed quantity is e times the probability that a pair of molecules will collide with a relative speed between
c and c + dc. Integrating (8.3) over all speeds leads immediately to (8.2).
Substituting (8.3) into (8.1) the reaction rate is
E,
) 112
(8.4)
which is precisely the form of the rate law for the elementary step being
studied; the bracketed term is the rate constant. Here nd 2 is the collision
(1)
(2)
Chap. 8
236
Single-Collision Chemistry
cross section which measures how close the centers of mass of A and
must pass in order for collision to occur. The factor p() accounts for
fraction of the collisions that lead to reaction. Both nd 2 and p( )
phenomenological parameters. Rather than treating each separately
define a cross section for reaction(3)
Be
the
are
we
dnA
[8
- ( f t = nAnBC nf-t(kT)3
1/2
IOO
o Q()e-
dkT
(8.5)
The reaction cross section is equivalent to the collision cross section for
hypothetical spherical molecules which react with each collision; in the
next section we indicate how Q() can be measured directly. The rate
constant is thus identified from (8.5)
k(T) = [
8
nf-t(kT)3
dkT
(8.6)
The d that measures the reaction cross section may be substantially different from a d
determined from gas-phase transport experiments. Ion-molecule and alkali metalhalogen reactions often have cross sections substantially larger than what would be
expected from analysis of transport data for the corresponding species. As we have defined Q(E), it is a function of molecular energy; however, experimental data usually
present Q(E), a function of molar energy. At 300 K the mean collision per colliding pair
is ~6 x 10- 21 J (equivalent to 0.04 eV or, in molar units, to 4 kJ mol-I).
Sec. 8.3
237
+ BC ---' AB + C
For a description of the apparatus necessary for production, detection, and analysis
see J. M. Parson and Y. T. Lee, J. Chern. Phys. 56,4658 (1972); Y. T. Lee, J. D. McDonald, P. R. Le Breton, and D. R. Herschbach, Rev. Sci. Insfr. 40,1402 (1969).
238
Chap. 8
..
Colhsion regIon
Source of A
Single-Collision Chemistry
Detector
~
of volume Vcoll
/;/;;
/V
' "\
/::(~~attered
r---L.~~~~i~~li~ 19 :lIision products
Beam of
intensity IA
---""_--'
Fig. 8.1. Schematic diagram of a molecular beam experiment. The detector is movable; it
subtends the solid angle dQ = sin (J dO drp.
react per unit time per Be molecule in the collision volume so that the total
rate of product formation is
d[Prod]
dt
[hQ(E)]NBC
(8.8)
N
VuexCu) du
Here the fraction of molecules in the specified velocity domain, the analog
to exCu) du, is 1 since the beam is velocity-selected. The approach speed
is now Crel> not u, and the beam intensity is [BC, not G(u) duo Thus, the
flux is
II
= ~Q(E)[BCh
e re1
(8.9)
Sec. 8.3
239
240
Chap. 8
Single-Collision Chemistry
L L fA(i)(T)f~8(T)Q(E, i,j)
i
(8.11 )
where fA(i)(T) is the fraction of A molecules in the internal state i at temperature T. From statistical thermodynamics this fraction is rfJ<i/qA,
where qA is the partition function for the internal degrees of freedom of
A. (5) With this result and its analog for the fraction of BC molecules, the
average cross section can be computed from (8.11); it is this average which
is substituted in (8.6) to calculate a rate constant.
If the colliding molecules can produce a variety of collision products,
there are separate reaction cross sections for each chemical transformation.
In the terminology of the field the individual pathways are called channels.
A channel is open if the total energy available, defined as the energy of the
separated reactants plus the collision energy, is greater than that of the
separated products. Even though a channel is open there may be no detectable product yield if the reaction cross section is sufficiently small. To
each channel there corresponds a rate constant for reaction in a system at
thermal equilibrium to be calculated using (8.11) and (8.6); the activation
energy is found from (8.7).
<
Eo
E> Eo
(8.12)
Even at this level of simplification the model cannot be correct. Not all
the collision energy is available to overcome the barrier. The fraction of
(0)
Sec. 8.4
Hard-Sphere Model
241
relative speed c
,.-- "'
'
, --l.
I
" '--
/
/
T
projection of c along line of
centers
=c
cos
e = c j 1- b2/(T2
Fig. 8.2. Diagram for determining line-of-centers velocity for hard spheres which do not
collide head-on; the impact parameter is b.
Since t(b) > to for reaction to occur, we can solve for b and find that, for
reactive collisions, b is less than a critical value be(t)
Collisions which occur with impact parameter b have their loci on a cylinder
of radius b. Thus, the collision cross section for an impact parameter
between band b + db is the annular area 2nb db. The fraction of collisions
which occur with this range of impact parameters is therefore 2nb db
divided by the total collision cross section na 2 Since reaction only occurs
if b < bit), the fraction of reactive collisions is
Chap. 8
242
Single-Collision Chemistry
n~f-t
- EO/E) dE
<0
(8.14)
In SI, the units of k(T) are m 3 molecule- 1 sec-I; to convert to the more
commonly used units liter mole- 1 sec- 1 multiply (8.14) by 103 No. The
activation energy can be computed from (8.7); it is
(8.15)
which, as expected, varies slowly with temperature. The activation and
threshold energies are not the same, which is generally the case. However,
the change in activation energy predicted by this model is ,....,2 kJ mol-1
for a temperature change of 500 K. Experimental determinations of Ea
are rarely both sufficiently accurate and made over a large-enough range of
temperature to render such variation observable.
The preexponential factor in the line-of-centers model is proportional
to T1!2 which is a direct consequence of the cross-section function (8.13).
With a different, but less readily rationalized, model for Q(E), a constant
preexponential factor is predicted (Problem 8.18).
Both Eo and ()" are parameters of the model. It is not possible to determine Eo from the properties of the individual reactants. However, ()" can be
estimated from the molecular diameters of isolated reactants using transport,
spectroscopic, or X-ray diffraction data. Thus, the preexponential factor in the
Arrhenius expression can be computed. From (4.42), (8.14), and (8.15) we find
8e )112
A= ( - - n()"2
nf3f-t
(8.16)
CI + H.
0.80
1.6
12
0.035
9.4
0.058
400
200
0.35
10
7
14
85
12
Experiment
900
920
1400
480
340
330
47
59
110
47
59
26
65
1070
Calculated
from (8.16)
~
a
~
w
'r~
..,~
:c...
00
Reference
a D. R. Herschbach, H. S. Johnston, K. S. Pitzer, and R. E. Powell, J. Chern. Phys. 25, 736 (1956), where original references are given.
b V. N. Kondratiev, Rate Constants of Gas Phase Reactions, translation by L. J. Hoitschlag (Washington: National Bureau of Standards, 1972),
where original references are given.
CCls + CI.
HCI + CI
H. + N.H.
NH. + NH.
HC] + CCl.
2Br+Ar
3.5
2
8
29
16
138
84
18
CI. + O.
CO + CI.
HI +1
10.5
43.5
132
36
102.5
NO. + O.
N0 2F + F
NO + CO 2
FCIO. + F
2NO + CI 2
Activation energy
(kJ mol-i)
H + CCl.
Br2 + Ar
CI + CCl,
H + 12
H + N.H,
H + N,H,
NO +0 8
N02 + F2
NO. + CO
F. + CIO.
2NOCI
2CIO
COCl + CI
Reaction
244
Chap. 8
Single-Collision Chemistry
the model is poor. Calculated values of A all fall in the fiftyfold range,
3 X 1013-1.5 X 1015 cm3 mole-1 sec-I, in disagreement with experiment, a
result that cannot be ascribed to errors in the estimate of cr. A twofold
error would only change A by a factor of 4 and the actual discrepancies
are usually much larger. In one case (H + N 2H 4 ), two reaction channels
with significantly different A-factors exist; collision theory makes little
distinction between them. In all but one case the calculated value of A
is much larger than that found experimentally, indicating that (8.l3) generally overestimates reaction cross sections.
Some of the defects of the hard-sphere model for the reaction cross
section are apparent. It assumes that all the line-of-centers kinetic energy
can be used for overcoming the threshold. It does not account for the
dependence of Q(E, i, j) on the internal states of the reactants. It does not
consider any effect due to molecular structure or to the details of the collision process. Improvement requires either direct experimental measurement of Q(E, i, j) or equivalently, a calculation which takes into account
the interaction potential between colliding molecules in specific internal states.
(7)
For an extension of the discussion in this and the next section see R. Wolfgang,
Ace. Chern. Res. 2, 248 (1969).
This term denotes that electronic energy is released in the reaction, i.e., the products are
more stable than the reactants. It is the equivalent, in discussions of chemical dynamics,
to the familiar word exothermic which is used to indicate that heat is evolved in a
reaction at constant temperature.
Sec. 8.5
Ion-Molecule Systems
245
0.5
0.2
0.1
0.05
Q(E) (nm 2 )
0.02
0.01
0.005
/
/
~
r
/J
co+ +
246
Chap. 8
Single-Collision Chemistry
Fig. 8.4. Effect of collision energy and impact parameter on trajectories when there are
long-range attractive forces and no threshold for reaction. (a) Variation of energy at constant impact parameter; (b) variation of impact parameter at constant collision energy.
illustrated in Fig. 8.4. When the impact parameter is large, only reactants
with low collision energy are deflected sufficiently to spiral in toward one
another and react. As the collision energy increases, the relative speed
increases and the trajectory is less curved. As E increases further, there is
less deflection and the ion and molecule no longer pass close enough for
reaction to take place. The qualitative picture is the same as that illustrated
in Fig. 2.9 for interaction via a Lennard-Jones potential; at large impact
parameters significant deflection only occurs for small collision energies.
The effect of varying E at constant b is shown in Fig. 8.4a. There is a critical
energy Ee(b) above which reaction cannot occur; as b increases Ee must
drop. The effect of varying b at constant E is shown in Fig. 8.4b. It is
complementary. At any E there is a critical impact parameter be(E) above
which no reaction can take place. As E increases, be must drop until it
becomes comparable with the "hard-core" distance (j which is determined
by the short-range repulsive forces between the reactant species. At this
and higher energies the repulsive forces are presumably most important
for determining reaction dynamics.
The existence of the critical impact parameter is often discussed in
terms of the so-called centrifugal barrier.(9) As particles spiral in toward
one another, the energy available for radial motion is less than the initial
collision energy. Conservation of angular momentum of two colliding
particles requires that there be nonzero centrifugal kinetic energy (that
part of the kinetic energy which accounts for motion perpendicular to the
line of centers). As the interparticle separation decreases so does the moment
of inertia; the centrifugal kinetic energy increases to conserve the angular
momentum.
For a pair with reduced mass fl, collision energy Eo, impact parameter
(9)
For a quantitative treatment of the relevant collision dynamics see R. E. Weston, Jr.,
and H. A. Schwartz, Chemical Kinetics (Englewood Cliffs, N.J.: Prentice-Hall, 1972),
Chapter 3.
Sec. 8.5
Ion-Molecule Systems
247
(8.18)
Here E ob2jR2 is the centrifugal kinetic energy. The radial speed dRjdt
is therefore governed by an effective potential E ob 2 jR2 + VCR). The first
term (the centrifugal barrier) is always positive and the second term is
negative for R large enough for particles to attract one another. Again
referring to Fig. 8.4b, which illustrates the effect of varying b at fixed Eo,
we note that as b increases, the distance of closest approach Ro (where
dRjdt = 0) also increases. When b is large, particles do not "collide";
instead they bounce off the centrifugal barrier and reaction, involving
approach to within distances comparable to chemical bond lengths, may
not occur. As b decreases, so does the centrifugal barrier; finally particles
may surmount it and penetrate to the repulsive wall. The critical impact
parameter bcCEo) is the maximum value of b for which such penetration
occurs.
For the Ar+ + D2 reaction the cross section is very large at low collision energies. Even for b < be(E), not all collisions are reactive; thus
for a given energy an upper bound to Q(E) is nb e2(E). The maximum
measured value of Q is '"'-'0.5 nm 2, which occurs at the lowest energy,
'"'-' 10 kJ mol- 1 . Thus, at this energy be is greater than 0.4 nm while the hardcore diameter, from Table 2.6, is '"'-'0.3 nm. The observed cross sections
are consistent with a model in which long-range attractive forces dominate
dynamics at low collision energy. Both the data and the model illustrate
an important generalization: the effective range of interaction decreases
as the collision energy increases.
Naturally if the products which are formed have lower bond energies, dissociation
occurs at smaller collision energies.
Chap. 8
248
Single-Collision Chemistry
upon the initial energy. We shall investigate this point when we study a
model for the mechanics of reaction dynamics in Chapter 10.
Thus far we have considered the barrierless, exoergic Ar+ + D2
reaction. To illustrate the effect of a potential barrier the reaction cross
section for
is included in Fig. 8.3. The threshold for reaction is ,-...,40 kJ mol-l. The
cross section is zero if the collision energy is less than the threshold value;
it then rises rapidly until the inevitable decrease occurs at high energy.
The low-energy cross sections for reactions with barriers are always significantly less than for those without barriers. At higher energies they are
comparable. At yet higher energies the cross section is determined by the
repulsive, short-range forces which are similar for all molecules.
+ K----" KI + CH 3,
LlE= -IOOkJmol- l
The cross section is plotted in Fig. 8.5. The threshold is small (;5 3 kJ)
and the cross sections large, comparable to those in the low-energy domain
of the Ar+ + D2 system (Fig. 8.3). The studies of the K + CH3I reaction
were not continued to energies where there is a precipitous drop in Q(E).
Even larger low-energy cross sections have been observed. Studies
of reactions of alkali metals with halogens are even more dramatic, as
long as there is no barrier. For the K + Br2 system a cross section of
,-...,2 nm 2 has been measured(ll); assuming each collision to be reactive,
(11)
Sec. 8.6
Atom-Molecule Systems
1.0
0.3
0.1
249
0.03
Q(E)(nm2)
0.01
+ O2
TO
T + O2
+0
~T +0 + 0
0.003
0.001
0.0003
Exchange
Threshold
0.0001
5
10
20
50
Dissociation
Threshold
100 200
Ecoll(kJ mol-I)
Fig. 8.5. Reaction cross section, Q(E), for three atom-molecule reactions as a function of
collision energy. [Data for K + CRaI reaction from M. E. Gersh and R. B. Bernstein,
J. Chern. Phys. 55, 4661 (1971). Data for T + D. reactions from M. Karplus, R. N. Porter, and R. D. Sharma, J. Chern. Phys. 45, 3871 (1966).]
reaction occurs at impact parameters of ;;:;0.8 nm, far larger than any
sensible "hard-core" distance. A reasonable chemical explanation of such
enormous cross sections may be based on the low ionization potentials
of alkali metals and the high electron affinities of halogens. Because of
this complementary relationship at distances of ;S 1 nm electron transfer
is highly probable, producing two charged species which are strongly
attracted to each other; reaction is then very likely. The mechanism has
been dubbed "harpooning" because of the following picturesque analogy.
At a fairly large distance the alkali metal "tosses out" an electron which
acts as a harpoon and attaches itself to the halogen; the ionic attraction
then provides the line for the whaler (alkali metal) to haul in the whale
(halogen). (12)
(1')
As is the case in small-boat whaling the distinction between the hauler and the hauled
is fuzzy. In addition, if the whale is powerful enough (large collision energies) it can
break the line and escape. In that event the whale carries the harpoon with it; here the
electron is transferred back (unless the collision energies are huge).
Chap. 8
250
Single-Collision Chemistry
+D
----'-TD
+ D,
is available; the results are included in Fig. 8.5. This cross-section function,
when used to compute rate constants via (8.6), led to results in good agreement with experimentally determined Arrhenius parameters. (13) The threshold is ,.....,,30 kJ mol-1 and the qualitative similarity with the C+ + D2
system (Fig. 8.3) is apparent. The decomposition cross section for the
reaction
11E = +437 kJ mol- 1
T + D2 -, T + D + D,
was also computed and is included in Fig. 8.5. Note that it does not show
any significant decrease at high energy, which is hardly surprising. If the
collision energy is large enough fragmentation and/or ionization is the
almost-certain outcome of any collision.
In the examples considered so far collision led to one of three results:
inelastic energy transfer, reaction, or decomposition. In more complicated
systems a bewildering variety of reactions becomes possible at large collision energy.u 4 ) The reactions of hydrogen atoms with propane provide
another good example. (15) These experiments were done using tritium
produced by neutron bombardment of 3He via the nuclear reaction
3He
+ n----'-p + T
Eo;S 50 kJ mol- 1
(13)
(14)
(15)
(16)
(8.19)
M. Karplus, R. N. Porter, and R. D. Sharma, J. Chern. Phys. 43, 3259 (1965); 45,
3871 (1966).
The K + CHaI system shows such behavior at higher collision energies. Possible
reactions have been cataloged by M. E. Gersh and R. B. Bernstein, J. Chern. Phys.
56,6131 (1972). The channel forming K+ + CHaI- followed by dissociation of the negative ion has been observed by A. M. C. Moutinho, J. A. Aten, and J. Los, Chern. Phys.
5, 84 (1974).
A. H. Rosenberg and R. Wolfgang, J. Chern. Phys. 41, 2159 (1964).
For a discussion of hot-atom chemistry see R. Wolfgang, Sci. Arner. 214(1), 82 (1966).
Sec. 8.6
251
Atom-Mdecule Systems
Table 8.2. Relative Rate Constants for the Abstraction, Replacement, and Double
Replacement Reactions in the T + C3HS Systetrt'
Temperature
(K)
Abstraction
(8.19)
Replacement
(8.20)
Double replacement
(8.21)
300
8 x lO- s
7 X 10-16
6x lO- N
600
6 X 10-4
8 X 10-8
I X 10-86
1000
2 X 10-'
1 X 10-'
4x 10-.2
(8.20)
(8.21 )
is one example, occur. The collision energy is then so large that there is
partial fragmentation. The generalization is obvious-as collision energy
increases more reaction channels are open. Reaction (8.21) need not occur
in a single step. If vibration ally excited C 2H 5T or CHaT, formed by replacement as in (8.20), has sufficient vibrational energy, it will undergo secondary
decomposition.
At thermal energies only abstraction is observable. From (8.6) the
quantity PEo exp( - pEo) provides a measure of relative rate constants if
we assume that the shapes of the cross-section functions near threshold
are not too vastly different. Values of PEo exp( -PEo) are given in Table
8.2 for the reactions (8.19)-(8.21). Only at 1000 K might replacement be
detectable; double replacement can never be an important thermal pathway.
Thus collision chemistry permits the study of reactions inaccessible by means
of ordinary thermal kinetic techniques.
252
Chap. 8
Single-Collision Chemistry
+ D2 ----' ArD+ + D
which we discussed as a prototype barrierless ion-molecule system, occurs at large impact parameter and proceeds via a direct, forward-peaked
mechanism. The reaction
D
+ Cl 2----' DCl + Cl
is favored by small impact parameters; it proceeds via a rebound mechanism. The third example
o + Br 2 ----' BrO + Br
is an indirect reaction; a Br 20 complex forms and survives for at least a
full rotational period (;:::;5 x 10-12 sec) before decomposing. The scattering
pattern is quite different in each case and provides a distinctive signature.
Molecular beam experiments are performed in a laboratory frame of
reference but the chemically interesting events take place with respect
to the center of mass of the colliding species. In order to interpret the
data, differential cross sections measured in the laboratory (LAB) coordinate system must be transformed to reflect events which took place in
the center-of-mass (CM) coordinate system. To effect this transformation
the invariant motion of the center of mass must be subtracted from the
scattering data obtained in the LAB system. A simple example which
illustrates the difference between LAB and CM kinematics is shown, for an
elastic collision, in Fig. 8.6. In CM the particles always move directly toward
one another before interaction and directly apart afterwards. This condition
is a consequence of momentum conservation in a system with a stationary
center of mass. The interaction causes each particle to be deflected through
Sec. 8.7
LAB
--VI
{v,
eM
,\2
Vc
253
~
VI
(a)
.,
.....
Uz
\.,,'"
.ul
(b)
Fig. 8.6. Comparison of collision kinematics in LAB and CM coordinate systems. Unprimed
and primed vectors refer to velocities before and after collision. In this example
M, = 2.7M2' (a) In LAB the center of mass, marked., moves with constant velocity
Vc throughout the collision. The collision geometry is for crossed beams at 90. (b) In CM
the center of mass is stationary. Conservation of momentum requires that relative motion
be linear before and after interaction. Each particle is deflected through an angle cpo
an angle rp. In LAB the situation is vastly different. The heavy partide,
which carries most of the center-of-mass momentum, is only slightly
deflected while the trajectory of the lighter particle is more significantly
perturbed. Similar diagrams can be constructed if collisions are inelastic
or reactive. The relationship between LAB and eM is more complicated
but, as long as the internal states of both reactants and products are known,
the correspondence can be made. (17)
A number of experimental limitations introduce ambiguities into the
determination of eM scattering patterns from LAB data. Molecules in
the incident beams have a range of velocities. There is therefore a distribution in initial center-of-mass and relative velocities. Thus a particular
deflection angle in LAB is correlated with a range of deflection angles
in eM. The uncertainty can be reduced by improving velocity selection in
the beams. Another source of ambiguity is that, in general, the scattering
pattern of only one collision product is determined which, unless the internal
energy of the product is measured, is insufficient to precisely characterize
the kinematic problem. A final difficulty is that molecules in the incident
beams are usually produced in a range of rotational states, which further
blurs the construction of the eM scattering pattern from LAB data.
(17)
Chap. 8
254
Single-Collision Chemistry
1.0
Relative
Intensity 0.5
O.OL-....::::..--~-...1---===O"'_-4
+5
+10
LAB angle
Ar+
Fig. 8.7. The angular distribution of ArD+ scattered in LAB due to the reaction
Ar+ + D2 ~ ArD+ + D. The Ar+ and D2 are incident at 0" and 90. The collision energy
is 7.8 kJ mol- 1 . [Adapted from Z. Herman, J. Kerstetter, T. Rose, and R. Wolfgang, Disc.
Faraday Soc. 44, 123 (1967).]
+ D2
System
The reaction
Ar+
+ D2~ArD+ + D
has been extensively studied using molecular beams. The scattering pattern
has been deduced by monitoring the ionic product ArD+. Measurements
have been carried out for collision energies from 7.8 to 880 kJ mol-I, an
energy range over which the reaction cross section drops by a factor of
about 50 (Fig. 8.3).(18)
The LAB angular distribution of ArD+ at the lowest collision energy
is shown in Fig. 8.7. All the ArD+ is scattered at angles close to the direction
of the incident Ar+ beam, which is expected since the center-of-mass momentum does not differ greatly from the momentum of the heavy incident
(Ar+) or scattered (ArD+) particles. By transforming these data to eM a
portrait of the reaction dynamics can be constructed.
The eM angular distributions of ArD+ determined for collision energies
of 7.8 and 440 kJ mol- 1 are compared in Fig. 8.8. In both cases the product
is predominantly forward scattered, i.e., it is moving in roughly the same
direction as the Ar+ incident beam, which reflects the chemical process
since the scattering pattern refers to eM, not LAB. The results reflect
(18)
Z. Herman, J. Kerstetter, T. Rose, and R. Wolfgang, Disc. Faraday Soc. 44, 123 (1967);
M. Chiang, E. A. Gisiason, B. H. Mahan, C. W. Tsao, and A. S. Werner, J. Chern.
Phys. 52, 2698 (1970).
Sec. 8.8
Direct, Forward-Peaked-Ar+
+ Da
255
System
center of moss
90
I
I
center of moss
I
I
I
I
90
(a)
(b)
Fig. 8.8. Polar diagrams showing angular distribution in CM of ArD+ formed in the
Ar+ + D. reaction at collision energies of (a) 7.8 kJ mol- l and (b) 440 kJ mol-l. The radius vector is proportional to ArD+ intensity. The diagrams are not drawn to the same scale.
[Derived from: (a) Z. Herman, J. Kerstetter, T. Rose, and R. Wolfgang, Disc. Faraday
Soc. 44, 123 (1967); (b) M. Chiang, E. A. Gislason, B. H. Mahan, C. W. Tsao, and A.
S. Werner, J. Chern. Phys. 52, 2698 (1970).)
o8
after reaction
(a)
before reaction
after reaction
(b)
Fig. 8.9. Model for spectator stripping. (a) Idealized: trajectories are not deflected due to
intermolecular forces. Product ArD+ is scattered at 0. (b) Effect of long-range forces.
Product ArD+ is forward scattered.
256
Chap. 8
Single-Collision Chemistry
impact parameter one D atom is snatched from the molecule while the other
is a passive spectator and continues along its original path with velocity
unaffected. The process is analogous to the parlor trick of snatching a
tablecloth out from underneath a set of dishes. If the cloth is pulled abruptly
the dishes do not move; the cloth goes with the prankster and the dishes
remain behind.
If the stripping model described the reaction precisely the Ar D+ would
appear only at 0, the direction of the incident Ar+. Real trajectories are
curved due to the forces between Ar+ and D 2. At low collision energy the
curvature is more pronounced (Fig. 8.4a) and so the ArD+ is considerably
deflected as shown in Fig. 8.9b. (19) At higher collision energies kinematics
always swamp effects due to longer range, weaker forces, as illustrated in
Fig. 2.9 and emphasized in the discussion of the centrifugal barrier (Section
8.5). The experimental consequence for the Ar+ + D2 system is that it
more closely approximates ideal stripping; as a consequence the forward
scatter is more pronounced.
Stripping, by its very nature, is a large impact parameter process.
At small impact parameters the reactants collide head-on; thus the Ar+
rebounds and is back scattered in the eM coordinate system either taking
a D ~tom with it or not, depending upon whether the collision is reactive
or nonreactive. Thus some back-scattered products are always expected,
(19)
There is a correspondence in the tablecloth-dish analogy. The less abrupt the tug on the
cloth, the more the dishes are perturbed. The reader is cautioned that expertise at the
trick comes at the expense of many broken dishes.
Sec. 8.9
Recoil-The D
+ ct.
System
257
center of mass
Fig. 8.10. Angular distribution in CM of DCl produced in the D + Cl. reaction. The radius vector is proportional to the intensity of DCI. The intensity at angles less than 60
is very small. [Derived from D. R. Herschbach, Disc. Faraday Soc. 55, 233 (1973).]
8.9. Recoil-The D
+ Cl
System
+ Cl
---->.
DC!
+ Cl
contrasts sharply with that in the Ar+ + D2 reaction. As shown in Fig. 8.10,
the DCl is predominantly back scattered which indicates that small impact
parameter recoil collisions favor reaction, (20) an observation consistent
Chap. 8
258
Single-Collision Chemistry
60
/
I---f
1000
m 8- 1
Fig. 8.11. Contour map of the angle-velocity flux distributions for DCl in the D + Cl.
reaction. The radius vector is proportional to the velocity of the Cl. Contours are drawn
at relative intensities of 0.1, 0.5, 0.8, and 1.0. [Adapted from D. R. Herschbach, Disc.
Faraday Soc. 55, 233 (1973).]
the collision energy of the reactants. (21) Referring to Fig. 8.11, most of the
product DCI appears at velocities between 1200 and 2000 m sec- 1 which
corresponds to energies between ,,-,50 and "-'150 kJ mol-I. The extra
kinetic energy acquired by the products derives from the large exoergicity
of the reaction, "-' 190 kJ mol-I. On the average "-'40% of the available
energy appears as product recoil energy; the remainder must reside as
internal energy (vibration, rotation) of DCI.
These observations can be used to determine the preferred reaction
geometry. As the D and Cl 2 approach they are decelerated due to repulsive
forces which create an activation barrier. The collision energy being greater
than the threshold energy, a D-CI bond forms with great energy release.
The products then recoil. In Fig. 8.12 two extreme geometries are shown.
As the D-CI bond forms the CI-CI bond is stretched and broken. Neglecting the mass of the D atom and the initial collision energy (which are
both relatively small), the CI atoms must recoil along the direction of the
original CI-CI bond in order to conserve momentum (the D atom is now
attached to one of the Cl atoms). The perpendicular geometry therefore
implies sideways scattering (peaks near 90) while the collinear geometry
is consistent with back scattering. Thus, the reaction probability is greatest
for collinear (or nearly collinear) collisions. (22)
The final energyinCM is E = (1/2)(MDClV~Cl + MCi V~l). Since momentum conservation requires that MDClVDCl + MClVCl = 0, Vel and thereby E can be calculated
from a measurement of VDCI.
( ) A rationalization, based upon consideration of the molecular orbitals involved in the
rearrangement process, is given by J. D. McDonald, P. R. Le Breton, Y. T. Lee, and D. R.
Herschbach, J. Chern. Phys. 56, 769 (1972).
(21)
Sec. 8.10
Collision Complexes-The 0
+ Bra
System
259
energy releas,e
- - - - upon reaction
4l \
1"'-;:"
cCb <=x=)
I
CI 2
~~
trajectories after
reaction
2:J
DCI
DCI
+ Br2
+ Cia
System
o + Br2~ BrO + Br
is shown in Fig. 8.13.(23) Unlike the reactions discussed in the two previous
sections the scattering pattern is symmetrical; the distribution is strongly
peaked at both 0 and 180, a result that cannot be understood in terms of
either the stripping or the recoil mechanism. However, if the reactants
form a collision complex a completely different picture emerges as shown
in Fig. 8.14. The reactants collide with an impact parameter, b (Fig.8.14a).
If the collision is "sticky" some of the relative momentum is converted into
angular momentum of the complex (Fig. 8.14b) so that the three atoms
rotate about the center of mass of the system. If the complex survives for
at least one full rotation before decomposing, all scattering directions in
the collision plane are equally probable, which accounts for the forwardbackward symmetry in a CM scattering pattern. The reason that the
measured angular distribution is not uniform can be understood from the
following considerations. Collisions with impact parameter b are cylindrically distributed about the z-axis of Fig. 8.14. Thus the full scattering
pattern is found by superposing the pattern for each collision plane as
indicated in Fig. 8.l4c. The axis of the resulting sphere is oriented along
(23)
Chap. 8
260
Single-Collision Chemistry
900
Fig. 8.13. Angular distribution in eM of BrO produced in the Br 2 + 0 reaction. The radius
vector is proportional to the BrO intensity. [Derived from D. D. Parish and D. R. Herschbach, J. Am. Chern. Soc. 95,6133 (1973).]
x
(0 )
b
Br2
---- 19
7------------
--- - z
-_/..-o
x (axis
of rotation
at complex)
(b)
----~--~~~--+-------~z
(el
iE--+---::' - - - - - - - ,
Fig. 8.14. Schematic diagram showing details in a collision typical of complex formation.
[Adapted from J. P. Toennies, Ber. Bunsenges. 72, 927 (1968).]
Sec. B.11
261
the direction of the relative velocity veClOr. All 0 and 180 product trajectories are superposed since these angles are the poles of the scattering sphere.
At other angles the trajectories are distributed on the parallels of latitude.
Thus as a detector is moved from 0 to 180 in a particular plane, observed
CM intensity is symmetric with its minimum at the equator (the 90 line).
The angular distribution of BrO shown in Fig. 8.13 is precisely what
is expected if complex formation has taken place. It is symmetric and
strongly peaked in both forward and backward directions. Such a pattern
only occurs if the complex survives at least one full rotational period
(;:;:;5 X 1O~12 sec) before decomposing. Since vibrational periods are typically ,....." 1O~13 sec, the Br20 complex is long-lived on a molecular time scale.
In this example vibrationally unexcited Br20 is stable relative to Br2 and O.
As is true of all cases in which complex formation has been found to occur,
there are strong long-range attractive forces (and thus no activation barrier)
which would lead to the formation of a stable molecule, radical, or ion if
the excess energy could be removed by collisional deactivation. Energy
removal is not possible given the conditions of a beam experiment. The
excess energy remains trapped m the vibrational modes of the complex,
being transferred between them until enough is concentrated in one mode
at which point decomposition occurs. Since the overall reaction
+H
262
Chap. 8
Single-Collision Chemistry
system is dramatic. (24) The reactive cross section for ground-state HCI
molecules is ,,-,0.0015 nm 2, less than 1% of that in the K + CHaI system
(see Section 8.6 and Fig. 8.5) which indicates there is an activation barrier,
estimated to be "-' 10 kJ mol-I. Reaction only occurs at small impact
parameters; a recoil mechanism is most likely. One quantum of vibration,
in this case ,,-,36 kJ mol-t, provides more than enough energy to surmount
the barrier and the cross section increases enormously to ,,-,0.2 nm 2, comparable to that found for the reaction K + CHaI. Now reaction takes place
at much larger impact parameters; the mechanism may well be totally
different.
Methyl iodide is a dipolar molecule; it can therefore be aligned in an
electric field. By passing a beam of CHaI molecules through such a field
a beam of oriented molecules is produced. This technique was used in a
study of the
Rb + CH3I ----'" RbI + CH 3
system. (25) Reaction was about four times more probable when the methyl
iodide molecules in the beam were selected such that the rubidium approached the iodine end of the molecule than if the alignment was such
that the rubidium approached the methyl end. The results are consistent
with the model in that electron transfer from alkali metal to halide is the
primary step.
A similar experiment was performed on the
K
+ CFaI----'" KI + CFa
system. (26) The results were strikingly different. Reaction is most likely
if the molecules in the beam are oriented so that the alkali atom approaches
the CFa end ofthe molecule. Presumably electron transfer is still the primary
step; since the F atom is far more electronegative than the I atom, it is
much more likely to participate in the harpooning step. However, the
C-I bond is weaker than the C-F bond (234 vs. 490 kJ mol-I) and the
further course of reaction seems governed by energy considerations. The
I atom is the one removed which simply requires that the charge in the
K +-CFaI - be rapidly redistributed after the electron transfer has taken place.
(24)
(251
(26)
263
Problems
Problems
8.1.
- 3RT/2
where
8.2.
Ea l1l
8.4.
8.5.
8.6.
Chap. 8
264
Single-Collision Chemistry
8.7.
8.8.
At 500 K the viscosity of H2(g) is 1.26 X 10-5 and that of I2(g) is 2.20 X 10-5
in SI units. Estimate the molecular size and then use collision theory to
obtain a value for the A-factor. The experimental value is ,....,4.3 X 1010
M-l sec-I.
_{O,y
B
+ H. ~ TH +
E
0.69
exp( -0.47y1.46),
H is roughly given by
<
Eo
E> Eo
where
y == (E -
Eo)/iJ,
Eo =
4.2 X 10-. 0 J,
iJ =
6.0 X 10-19 J,
B = 0.039 nm'
2"
Problems
265
Table 8.3
a'Q/ao alP
a2Q/ao alP
(nrn 2 rad-')
(deg)
(nrn" rad- 2 )
(deg)
(i)
(ii)
(i)
(ii)
0.0048
0.16
105
0.060
15
0.15
120
0.049
0.0054
30
0.14
0.0015
135
0.047
0.0061
45
0.13
0.0021
150
0.047
0.0066
60
0.10
0.0026
165
0.044
0.0068
75
0.079
0.0033
180
0.043
0.0074
90
0.071
0.0040
8.13. For reactive scattering of Na and (i) Br., (ii) CHsI at about the same collision energy the variations of differential cross sections with scattering
angles (CM) are [data derived from J. H. Birely, E. A. Entemann, R. R.
Herm, and K. R. Wilson, J. Chem. Phys. 51, 5461 (1969)] as given in
Table 8.3.
where q = (m;c' - m2c)/ M. If q is small show that cos 0 = l-q" sin" W/2Vi"'
where cos W = (Vi q)/I Vi II q I
266
Chap. 8
Single-Collision Chemistry
8.15. In idealized spectator stripping the atom left behind continues with its
velocity unaltered. The reaction Ar+ + D. ~ ArD+ + D is 145 kJ mol- i
exoergic.
(a) Use conservation of momentum and show that, in ideal stripping, the
relative translational energy after reaction is E' = 1OEo/21, where Eo is
the collision energy.
(b) Since excess collision energy must appear as internal energy of ArD+,
at what collision energy does the reactive cross section for stripping go to
zero? The binding energy for ArD+ is 370 kJ mol-i.
8.16. A "steric factor" is often used to account for discrepancies between observed A-factors and those computed using the hard-sphere model (S.16).
The steric factor / accommodates orientational requirements in reaction
so that the A-factor is A = Ao/with Ao given by (S.16).
(a) What are the steric factors for the reactions of Table 8.1?
(b) Comment, if you can, on possible reasons for differences between /
in the following pairs of reactions:
(i)
CI
+ CCI 4
H
H
(ii)
+ 12
CI + H2
(iii) H
CCI.
CI 2
+I
~ HCI + CI
~
HI
<
Eo
10 :::::: Eo
+ H. ~ DH + H+
+ H. ~ DH + H-
General References
267
General References
J. L. Kinsey, in MTP International Review of Science, Physical Chemistry, Ser. I, Vol. 9,
J. C. Polanyi, ed. (London: Butterworths, 1972), pp. 173-212.
K. J. Laidler, Theories of Chemical Reaction Rates (New York: McGraw-Hili, 1969),
pp. 182-204.
R. D. Levine and R. B. Bernstein, Molecular Reaction Dynamics (Oxford: Clarendon
Press, 1974), Chapter 6.
.
J. Ross, ed., Molecular Beams, Advances in Chemical Physics, Vol. 10 (New York: Interscience, 1966).
Ion-Molecule Reactions
P. F. Knewstubb, Mass Spectrometry and lon-Molecule Reactions (Cambridge: University
Press, 1969).
Elastic Scattering
R. J. Cross, Jr., Acc. Chern. Res. 8, 225 (1975).
Inelastic Scattering
G. A. Fish and F. F. Crim, Acc. Chern. Res. 10,73 (1977).
I. W. M. Smith, Acc. Chern. Res. 9, 161 (1976).
J. P. Toennies, Ann. Rev. Phys. Chern. 27, 225 (1976).
Reactive Scattering
D. R. Herschbach, Disc. Faraday Soc. 55, 223 (1973).
D. A. Micha, Acc. Chern. Res. 6, 138 (1973).
State-to-State Experiments
J. M. Farrar and Y. T. Lee, Ann. Rev. Phys. Chern. 25, 357 (1974).
270
Chap. 9
f~ (27tb db)p(b, E, i, j)
(9.1)
{~:
and be (E) = 0'(1 - Eo/)1/2. Substituting this expression into (9.1) leads
immediately to the result for the reaction cross section obtained previously
(8.13). For more realistic models the determination of the reaction probability is a formidable problem which requires, as a preliminary step, a
knowledge of the interaction potential (the potential-energy surface) of the
reactants as a function of the nuclear coordinates. Such a calculation
determines the electronic energy of the interacting molecules for all nuclear
configurations of interest. Once the potential-energy surface is known
the problem of computing the reaction probability can be tackled. This
requires applying time-dependent quantum mechanics to treat the problem
of nuclear motion in a known potential. The task is exceptionally difficult
and is one that has not been accomplished, even for the simplest reactions,
at the present time. (1)
(1)
Recent calculations on the H + H2 ~ Ha + H exchange reaction come close to carrying out this program; see A. Kupperman and G. C. Schatz, J. Chem. Phys. 62, 2502
(1975); A. B. Elkowitz and R. E. Wyatt, J. Chem. Phys. 62, 2504 (1975).
Sec. 9.2
Potential-Energy Sur/aces
271
+ BC----"AB + C
(9.2)
three parameters are needed, e.g., the distances A-B and B-C and the
A-B-C angle. (2) Correspondingly the surface is four-dimensional. If
one of these variables is held fixed, a three-dimensional projection can be
constructed. A series of such diagrams permits visualization of the whole
potential-energy surface.
Reactions that proceed via an intermediate which is more energetic
than either reactants or products, as in Fig. 4.8, have qualitatively similar
potential-energy surfaces. A typical slice at fixed A-B-C angle is shown in
perspective in Fig. 9.1 and as a contour map in Fig. 9.2. When rAB (the
A-B distance) is large, the interaction potential is that of the isolated BC
molecule; the point R represents the classical ground state of the reactants.
Similarly when rBC is large the potential is that of the AB molecule; P is
the classical ground state for products. Large values of both rAB and rBC
correspond to total dissociation into A + B + C; this region of the surface
is the plateau D.
Reaction is represented by motion from R to P on the surface. As
reaction occurs the system starts in the reactant valley, travels over a pass,
and descends into the product valley. Reaction only occurs if the relative
energy of the interacting molecules is sufficient to surmount the pass. For
thermal reactions the system must move along relatively low-energy paths
because, for surfaces like that of Fig. 9.1, the pass height is usually high
enough so that collisions which occur with enough relative energy are
highly improbable. Reaction thus samples the high-energy tail of the Boltzmann distribution and only collisions with energy close to the pass energy
are sufficiently probable to contribute significantly to reaction. (3) The
minimum energy path is shown in both Figs. 9.1 and 9.2. The section through
this path, known as the reaction profile, is shown in Fig. 9.3. It is similar
to the schematic diagram of the activation process shown in Fig. 4.8;
now, however, progress along the path is correlated with specific changes
in geometry. Reactants with collision energy less than the energy of the pass
(2)
(3)
Many other ways of specifying the ABC triangle are possible but this is the most convenient choice.
If the system strays too far from the pass the collision energy must increase in order for
reaction to be possible. At 300 K the fraction of possibly reactive collisions decreases
tenfold with each 6 kJ mol- 1 increase in the barrier.
Chap. 9
272
Potential
Energy
Fig. 9.1. Variation of potential energy with A-B and B-C distances for an ABC system in
which the A-B-C angle is held fixed. [Adapted from K. J. Laidler, Chemical Kinetics (New
York: McGraw-HilI, 1965).)
(A+B+CI
(A+BC) R
Fig. 9.2. Contour diagram for the potential energy of the ABC system of Fig. 9.1. The
dashed line shows the minimum energy path which is followed by a reaction under thermal
conditions.
Sec. 9.2
Potential-Energy Sur/aces
273
Activated complex
ABC
Potential
Energy
A+BC
dx
AB+C
Reaction coordinate
Fig. 9.3. Section through the minimum energy path of Fig. 9.2. There is a barrier to reaction
in either direction.
do not react. (4) In thermal chemical reaction, only trajectories close to the
reaction profile are important; all others require too much collision energy.
Thus most reaction paths pass close to the pass on the potential-energy
surface. Because of the chemical importance of the relatively low-energy
region of the ABC aggregate system, it is specially designated. The pass,
which is the peak of the reaction profile, is known as the transition state
or the activated complex. It bears, of course, no relation to the long-lived
collision complexes discussed in Section 8.10.
Changing the A-B-C angle affects both the potential-energy surface
and the reaction profile but only in those regions where the three particles
are close together. The far valleys and the plateau which describe the
isolated molecules and the separated atoms are not changed. Thus all
reaction profiles are alike at start and finish but their intermediate structure
differs. The pass energy depends upon the angle A-B-C as is illustrated
in Fig. 9.4, where the reaction profile for the exchange reaction
is plotted for different H-H-H angles. The lowest pass height in this system
occurs in the linear configuration. The activated complex or transition state
is identified as the configuration for which the ABC complex has the
minimum pass energy. It is important to recognize that the most probable
reaction paths pass through configurations close to that of the activated
(4)
This is not strictly correct. Molecules follow the laws of quantum, not classical, mechanics. As such they have zero point energy; the system does not precisely follow the potential-energy trough which slightly alters the collision energy necessary for surmounting the pass. It is also possible, but not likely, for reactants with collision energy somewhat less than the pass energy to tunnel through the barrier. These quantum effects
alter the threshold energy which is usually slightly less than the pass energy.
Chap. 9
274
Potential Energy
(kJ mol-I)
-0.05
r(nm) 0.05
Fig. 9.4. Reaction profiles for the Ha system as a function of HHH angle. The distance r
is the separation of the incoming or leaving H atom from the pass on the potential-energy
surface. Energies are measured with respect to the H + H, system. [Adapted from I. Shavitt, R. M. Stevens, F. L. Minn, and M. Karplus, J. Chern. Phys. 48, 2700 (1968).]
(0 )
>.
>.
0>
0>
Q)
Q)
'0
:;:
c
:;:
c
0
c..
c..
Q)
Q)
'0
Reaction Coordinate
Reaction Coordinate
Fig. 9.5. Minimum energy path for systems without barriers. (a) Typical profile for reaction with shallow well when complex formation does not occur. (b) Typical profile for
reaction with deep well where intermediate complex may form.
Sec. 9.2
Potential-Energy Surfaces
275
Chap. 9
276
+B
---->.
products
(9.3)
with a reaction profile like that of Fig. 9.3. Motion along the reaction
coordinate is specified by a position x and a reaction momentum p. By
P > 0 we mean that the reactants are moving toward each other; as they
approach the pass they may be considered to form an aggregate or a composite system (sometimes described as an AB complex). The transition from
independent reactants to aggregate system cannot be recognized to have
occurred at any particular point along the reaction coordinate. At small
separations (;50.5 nm) interaction is strong and the composite description
is appropriate. At large separations (;;:; 10 nm) interaction is weak and
describing the reactants as independent species is more reasonable.
The basic assumption of the theory is to consider motion along the
reaction coordinate separately from other motions on the potential-energy
surface. In addition to its translational, rotational, and vibrational modes
the AB aggregate has a reaction mode which describes the chemical transformation in terms of progress along the reaction coordinate. The larger
the value of the reaction momentum p, the more collision energy is available
for surmounting the pass. If the energy in the reaction mode is large enough,
the molecules traverse the pass and react; if not, they are reflected. (7)
Far away from the pass, the number of A-B composite systems, C(p, x),
with reaction coordinate between x and x + dx and reaction momentum
(6)
(7)
The development which folIows is adapted from H. Eyring, J. Walter, and G. E. Kimball, Quantum Chemistry (New York: John Wiley, 1944), pp. 299-311. For a totally
different development see K. J. Laidler, Theories of Chemical Reaction Rates (New York:
McGraw-HilI, 1969), pp. 45-53.
Again this is not quite true since the problem is one of quantum, not classical, mechanics. See footnote (4) in this chapter.
Sec. 9.3
277
where f1 is the reduced mass for motion along the reaction coordinate,
f1 = mAmB/(mA
mB)' The total number of systems approaching the
C(P,x)-df1 x
(9.4)
T. L. Hill, Introduction to Statistical Thermodynamics (Reading, Mass.: AddisonWesley, 1960), Chapters 4, 8, and 9. A summary of some important results of statistical thermodynamics is included in Appendix A.
278
Chap. 9
Cn(p, X) dp dx
NAB
(
qAB exp -
+ dp,
and
En
En)
(
kT exp -
is(9)
p2) dp dx
2flkT - h -
(9.6)
Combining (9.6) with (9.5) and summing over all degrees of freedom
except the reaction coordinate and momentum we find
C(p, x) dp dx
NANB:I:
(
qAqB qAB exp -
p2) dp dx
2flkT - h -
(9.7)
qiB
~ exp ( - :~ )
(9.8)
~)
;
p
( - - - - dp
NANB~-exp
qAqB
flh
(9.9)
2flkT
dNA
qiB
---=NAN
Bdt
qAqB
foo --x(p)exp
P
(P2-) dp
0
flh
2flkT
See Appendix A for a justification of this expression. The factor of h- 1 occurs because
the formulation is semiclassical. It is against the rules of quantum mechanics to treat
reaction coordinate and momentum separately.
(10) This is an excellent approximation in dilute gases. At 1 atm and 500 K the mean free
path is ~50 nm, much larger than the range of values of the reaction coordinate that
can be considered to correspond to the composite. At very low pressure the approximations used to derive (9.7) are no longer reasonable; under such conditions the concentration of composites is sensitive to the rate at which they traverse the pass. In
addition, depletion is significant and (9.5) is no longer a good approximation.
(9)
Sec. 9.3
279
(9.12)
where <x) is a mean transmission coefficient. Its value, determined by
comparing the computed and experimental A-factors, is generally between
0.1 and 0.9; in practice it is often assumed to be ,..".,0.5. As formulated,
k 2 (T) has units of molecules- l sec-I. By introducing appropriate factors
of V and No it can be converted into concentration units, cm 3 mol- l sec-I.
In terms of the molar partition function per unit volume,
(9.13)
(11)
Chap. 9
280
x(p)
.- /
I
I
i
I
I
I
I
p'f =
I
I
I
\...1
htJE'
vi
--
reaction momentum
Fig. 9.6. Behavior of transmission coefficient ,.;(p) as a function of reaction momentum for
classical and quantum mechanical models of reaction; a rectangular barrier is assumed.
which is the famous Eyring equation. The rate constant can be calculated
from the properties of the potential-energy surface by identifying the reaction coordinate. The pass height determines the Eyring activation energy
E~, here defined as the classical barrier to reaction; as such it does not include
the zero point vibrational energy of the activated complex. The shape of
the reaction profile determines ,,(p) and thus <,,); the remainder of the
surface near the pass determines Q1B'
Transmission
Function
(Arbitrary
units)
classical
quantum
reaction momentum
Fig. 9.7. Transmission function, p,.;(p) exp( - p2 /2{1kT), as a function of reaction momentum
for classical and quantum models of ,.;(p) plotted in Fig. 9.6. Note that oscillations in
,.;(p) are damped by the Boltzmann factor; E+ /kT is much greater than 1 corresponding to a
significant activation barrier.
Sec. 9.4
281
As was the case in the hard-sphere model for reaction rates (Section
8.4), c t is not easily computed and thus a complete calculation of k 2 (T)
is not generally possible. To eliminate c t and determine an expression for
the A-factor we write (9.14) as
k 2(T) = B(T) exp( -ctjkT)
B(T)e B
(9.15)
Again both A and Ea are temperature dependent but the variation is usually
too small to be detected experimentally.
The theory is obviously quite approximate even if <u), c t , and Qi.B
could all be computed. We began by assuming an equilibrium population
of A-B composite systems at points along the reaction profile far from the
pass; however, the problem is intrinsically nonequilibrium in nature. We
assumed that motion along the reaction coordinate was not coupled to
other degrees of freedom of the A-B composite, which is highly unlikely,
if only because the moments of inertia of the aggregate vary as the reaction
coordinate changes. We assumed the one-dimensional reaction coordinate
could be uniquely assigned even though molecular beam experiments
indicate that reaction dynamics may be much more complex. Even with
these limitations the theory is still a considerable improvement over the
hard-sphere model of Section 8.4. Its great achievement is that, through
consideration of the partition function Qh, plausible assumptions as to
the nature of the A-B interaction can be introduced and used to calculate,
in an admittedly approximate manner, either k 2 (T) or, more commonly,
the A-factor.
282
Chap. 9
surface, the properties of the activated complex are usually estimated. These
estimates then provide the data used to determine k2(T).
Here we treat two specific examples. First we consider the hard-sphere
model for chemical reaction and show that activated complex theory leads
to a k 2(T) identical with that calculated using kinetic theory arguments in
Section 8.4. We then consider the molecular example
H
+1
2 ----'
HI
+I
ge
(12)
(13)
(14)
We assume that only one electronic state of the complex is important in thermal reactions.
It might seem that this assumption of separability implies that knowledge of the saddle
point structure can be used to characterize the reactants and products. If each degree
of freedom were strictly independent then precise description of the saddle point
structure would specify the molecular properties of both reactants and products. Since
separability is only approximate, no such extrapolation is possible. In fact, the reaction
coordinate evolves from a molecular vibration in a manner unspecified by knowledge
of the saddle point region.
T. L. Hill, Introduction to Statistical Thermodynamics (Reading, Mass.: AddisonWesley, 1960), Chapters 4, 8, and 9; or see Appendix A.
Sec. 9.4
283
+ I2~ HI + I
(9.20)
Chap. 9
284
linear molecule); v, the vibrational frequencies; and (1, the symmetry number. If HI2 is assumed to be linear with interatomic distances 30% greater
than those in HI or 12, the following set of molecular parameters can be
obtained(16) :
H: g=2
12 : g = 1, (1 = 2, 1 = 7.5 X 10-45 kg m 2, v = 6.45 X 1012 sec-1
<x>
kT
h
Q~I.
QHQI.
which may be evaluated using (9.16) and the statistical mechanical expressions for the various partition functions, (17)
(16)
(11)
With a 30% bond expansion, the bond lengths in HI. are /(H-I) = 0.21 nm and /(1-1) =
0.35 nm. Calculation of the vibrational frequencies requires estimating the force constants
in the activated complex. In the series HF, HCI, HBr, and HI, the quantity [(frequency)
x (bond length)] is essentially constant. Assuming this to be generally applicable, the
vibrational frequency of HI (6.94 x 1013 sec- 1 ) leads to the estimate given for the stretching frequency of HI.. The bending frequency in HI. cannot be estimated from the
properties of HI. However, assuming that HI and HTe bonds are similar, a bending <L
force constant, k~, can be determined from the spectroscopic properties of H.Te (bend-.'
ing frequency 2.58 x 1013 sec-l, bond length 0.165 nm). Using this value for k~, the.
bending frequency for both linear and bent HI. may be calculated. Formulas.. relating
force constants to frequencies are given by G. Herzberg, Infrared and Raman Spectra
(Princeton, N.J.: Van Nestrand, 1945), pp. 168-175.
T. L. Hill, Introduction to Statistical Thermodynamics (Reading, Mass.: AddisonWesley, 1960), Chapters 4, 8, and 9; or see Appendix A.
Sec. 9.4
285
HI
+I
Calculated
180
90
a
250K
500K
750 K
0.055 (0.034)
0.055 (0.028)
0.073 (0.042)
3.04
2.73
2.95
(2.70)
(2.13)
(2.13)
Two possible geometries for the activated complex are considered. At each geometry the
first value is based upon a 30% bond expansion; the values in parentheses are based upon
no bond expansion. The experimental value for A in the temperature range 300-700 K is
2 x )014 em" mol- 1 sec-I.
2.05 X 10
<,,> ---=T=1/=214
(l (l -
e-310IT)
e-2560IT)
(9.21)
= 2,
= 1,
ABC
= 3.63 X 10-135 kg 3 m 6 ,
Here ABC is the product of the three principal moments of inertia. (18)
The difference is that one vibrational degree of freedom has been altered
to a rotational degree of freedom. By using statistical thermodynamics a
result analogous to (9.21) is found. The results of calculations assuming
linear and bent configurations for the activated complex are given in
Table 9.1 for three temperatures assuming both 30% and no bond expansion. A
of 1 was assumed so that the calculation should overestimate
the A-factor if the properties postulated for the HI2 complex are reasonable.
Some interesting points emerge. The calculation is insensitive to the
amount of bond expansion assumed. The temperature dependence of A
for the two geometries is quite small, which is a general feature so that
any temperature in the experimental range can be used in these calculations.
<,,>
(18)
The procedure for calculating the principal moments of inertia is given by I. N. Levine,
Molecular Spectroscopy (New York: John Wiley, 1975), pp. 197-203.
Chap. 9
286
Reaction
Experiment
+ 0 ~ NO. + O.
NO. + F. ~ NO.F + F
NO. + CO~ NO + CO.
Fa + CIO. ~ FCIO. + F
2NOCI ~ 2NO + Ci.
2CI0 ~ CI. + O.
COCI + CI ~ CO + CI.
NO
0.80
1.6
12
Collision
theory
Activated
complex theory
47
0.44
59
0.12
110
6.0
0.Q35
47
0.082
9.4
59
0.44
0.058
26
0.010
65
1.8
400
D. R. Herschbach, H. S. Johnston, K. S. Pitzer, and R. E. Powell, J. Chern. Phys. 25, 736 (1956).
The A-factors for the two geometries are very different. The bent activated
complex leads to a reasonable value; the linear complex leads to a gross
underestimate. In the bent geometry the calculated A-factor is 1.5 times
the experimental value, which is remarkable considering the approximate
nature of the calculation. In the linear geometry the calculated value is
30-40 times too small. Thus we can argue that this structure for the activated complex is very unlikely. (19) The calculation discriminates against
one geometry. It does not prove the other to be correct. Nonetheless given
a choice between a linear or nearly linear activated complex and one that
is strongly bent, there is reason to opt for the latter.
Similar calculations have been carried out to estimate the A-factor
for various reactions. The results, for some of the examples of Table 8.1,
are given in Table 9.2. The calculations did not account for bond-length
increases in the activated complex. Since expansion affects both the moments
of inertia and the vibrational frequencies in the same way, recalculation
would increase the predictions of the theory by factors of 3-5. Even without
this correction the predictions of activated complex theory are generally
within a factor of 10 of the experimental value. Only when the experimental
(19)
Sec. 9.5
287
A is very large are there gross discrepancies. However, there are other
difficulties. According to the theory as developed in Section 9.3, the mean
transmission coefficient is less than 1. Thus activated complex theory should
always overestimate k 2 (T), and therefore the A-factor as well. Even recognizing that the method of calculation used in making the predictions of
Table 9.2 leads to underestimates, the differences between calculation and
experiment are too large to be attributable to neglect of bond expansion in
the complex. Whether the differences are due to weaknesses in the theory
leading to (9.14) or to errors in the proposed structure of the activated
complex is unclear. In any case activated complex theory provides another
tool, albeit one which must be laced with a liberal dose of chemical intuition, to be used in studying a chemical reaction. It must be emphasized
that agreement between calculated and observed values of A does not
"prove" anything. As with alI other kinetic arguments disagreement indicates that the postulated mechanism or structure for the activated complex
is incorrect. A case in point is the H2 + 12 reaction. The A-factor calculated
assuming the Bodenstein mechanism
is 4 X 1013 cm 3 mol -1 sec-I, half the experimental value. (20) In spite of the
satisfactory agreement, there is other kinetic evidence that the reaction
does not proceed via the bimolecular mechanism.
+ BC-----"AB + C
where B can exist in a number of isotopic forms. Then, from (9.14), the
ratio of the rate constants for reaction involving different isotopes Band
B' is
Q(BC)
(9.22)
Q(B'C)
(20)
288
Chap. 9
The partition functions are determined by nuclear motion on the potentialenergy surfaces for the BC molecule or the ABC complex, and the activation
energy is the potential step required to traverse the lowest energy path on
the ABC surface. The phenomenon is a direct manifestation of quantum
mechanics. In the classical limit (massive species) the ratio is 1.
Simplification of (9.22) is based on the fact that the potential-energy
surface is the ground state for electronic motion in a charge distribution
determined by the nuclear configuration. Since isotopes have the same
nuclear charge, the potential-energy surface is unaffected by isotopic substitution. As a consequence the classical barrier height is isotope independent
and (9.22) becomes
k'
k
Q(BC)
Q(B'C)
<x')
<x)
(9.23)
Furthermore the shape of the surface in the reactant valley and at the
saddle point is not changed. The force constants for nuclear motion in
either region are invariants. Knowledge of the moments of inertia and
vibrational frequencies of one isotopic species allows computation of the
same quantities for any other. For a diatomic molecule like BCthe results
are particularly simple
I' _ (
)2 _
T - --;;- -
mB,(mB
mB(mB'
+ me)
+ me)
Similar, but more complicated, expressions can be found relating the properties of any molecular species.
Further simplification of (9.23) can be carried out in different ways.
In one, the quantities Q+ and Q are determined by a calculation similar
to the one used in Section 9.4 to compute the A-factor for the H + 12
reaction, leading to an expression involving masses, moments of inertia,
and vibrational frequencies. The other method depends upon the solution
of the vibrational secular equation in the reactant valley and at the saddle
point. (21) The result, for a linear transition state, is
k'
<x')
T - <x)
(21)
v'+
vr
r(BC)
r(B'C)
r 1(AB'C) [ r 2(AB'C)
r 1(ABC)
r 2 (ABC)
]2
(9.24)
Detailed analyses are given by J. Bigeleisen and M. Wolfsberg, Adv. Chem. Phys. 1,
15 (1968); L. Melander, Isotope Effects on Reaction Rates (New York: Ronald Press,
1960), Chapter 2.
Sec. 9.5
289
where
=--
2n
(-k
)1/2
-fl
where k is the curvature 0) and fl is the reduced mass for motion along
the reaction coordinate. Clearly vt, while having the dimensions of frequency, has no relation to the rate at which systems cross the pass; that
quantity is determined by (9.9).
After the potential-energy surface near the saddle point has been
constructed and the properties of the activated complex determined, all
factors in (9.24) except for the transmission ratio </(')/</() may be immediately computed. The transmission ratio may be calculated if highly
simplified models of the reaction profile are used. However, in most instances little error is introduced by assuming that </(')/</() = 1.(22) The
exceptions are for reactions in which isotopes of hydrogen are intimately
involved in the reaction process. While both the transmission ratio and the
partition function ratios of (9.24) approach 1 for massive reactants, the
transmission ratio tends to 1 more rapidly. Therefore its influence on the
isotope effect is relatively less significant, except at low temperature.
A system for which the effect of isotopic substitution has been thoroughly investigated is the reaction
Cl
+ H2 ----' HCl + H
k.[C1][iH jH]
~J
'
i, j = 1,2,3
Experimental values for the isotope effect knlk ij are plotted in Fig. 9.8.
Values computed from (9.24) are included in the figure. The parameters
(22)
Chap. 9
290
50~
20
kll
10
kij
.... .. ..
3.0
IH3H
IH2H
4.0
Fig. 9.8. Kinetic isotope effect for the reaction H. + CI ~ HCI + H. Solid lines calculated
using (9.24). [Adapted from R. E. Weston, Jr., Science 158, 332 (1967), where original references are given.]
required to evaluate (9.24) were calculated using a semiempirical potentialenergy surface; variation of the transmission ratio was taken into account.
Note that the kinetic isotope effect can be quite large; 3H 2 at the lower
temperature studied (273 K) reacts 35 times slower than IH 2 The calculated
values are in fair agreement with the data. The trends are correctly predicted
and numerical discrepancies are not large. It appears that deviations become
more severe at lower temperatures. Whether these are due to the weaknesses
of transition-state theory itself, the model used for the potential-energy
surface, or the way in which the transmission ratio has been estimated is
unclear.
Isotopic substitution in the Cl + H2 system leads to a "normal"
isotope effect, i.e., [k(light)]/ [k(heavy)] > 1. In other systems this ratio
is less than I-the so-called "inverse" isotope effect. The isotope effect is
determined by the difference between the shape of the potential-energy
surface at the reactant configuration and at the pass. Ignoring vibrational
excitation, the ratio (9.24) reflects differences in zero point energies (Problem
9.15). In the example considered the zero point energy of the activated
complex is less than that of the reactants-the activated complex is "floppy."
Sec. 9.6
291
If the activated complex were "constrained," as is the case when the isotopic
reactants are atoms, the opposite would be true; an "inverse" isotope effect
is expected. (23)
_ d[A]
dt
k* [A]
(9.25)
For products to form, the ethyl-carbonyl C-C bond must stretch and the
H atom on the carbonyl group must approach the ethyl group. Only if
sufficient internal energy is concentrated in a "critical mode" that leads
to reaction can decomposition occur. The total internal energy of a molecule
is distributed among all its internal modes. Only rarely is it concentrated
in a single vibrational mode. Thus, if the internal energy E is just greater
than the threshold energy Eo, few molecules will be collisionally activated
to a state for which sufficient energy is in the critical mode; the decomposition probability is thus small. As E increases, the likelihood of finding an
energy Eo in the critical mode also increases thus enhancing the decomposition probability.
(23)
K. J. Laidler, Chemical Kinetics, 2nd ed. (New York: McGraw-HilI, 1965), p. 97.
292
Chap. 9
= 4nd 2 (kTjnm)1!2
(9.27)
(9.28)
A factor of 2 is introduced because in the collision process either of the
A molecules might be excited.
While the population of low-lying states identified with A is not
altered by reaction, the population of the states identified with A*, i.e.,
those with sufficient energy to react, is affected. The total concentration
of A* molecules with energy between E and E dE is n~(E)P(E) dE, where
HE) is a depletion factor accounting for reaction. Since A *-A* collisions
are highly unlikely, they can be ignored. The total collision rate of A *
molecules of the specified energy is yn[n~(E)P(E) dE). Since the overwhelming
majority of A-A * collisions lead to deactivation, the total rate of destruction
of A * molecules of energy E is
(9.29)
(24)
We ignore any relation between collision frequency and the internal energy of the
products of the collision. Furthermore, the effect of the intermolecular potential is
ignored. To lift these restrictions a more general formulation of the problem, based
upon the "master equation" must be constructed. The master equation focuses upon
the transition probabilities botween molecular states. For a discussion and a general
review of unimolecular reactions see D. C. Tardy and B. S. Rabinovitch, Chern. Rev.
77, 369 (1977).
Sec. 9.7
293
The second term accounts for the effect of reaction; k(E) is the reaction
rate constant for molecules of energy E. Once a steady state is established
(9.28) and (9.29) are equal and we find
1
(9.30)
+ k(E)/yfJp
where the ideal gas law, n = fJp, has been used. At high pressure there is
little depletion of excited states due to reaction. At low pressure the effect
is important; molecules react too rapidly for the excited-state population
to be maintained by collisional activation. At low energy, where k(E) = 0,
there is also no depletion.
The total rate of product formation is found by integrating the second
term in (9.29) over all energies greater than Eo, the threshold for reaction
Note that at low pressure the rate constant is determined by the fraction
of activated molecules. If k(E) is assumed to be constant for E > Eo, (9.31)
reduces to the Lindemann expression (9.26).
Chap. 9
294
sel. (27) The modern picture, based, as is activated complex theory, (28) upon
identifying the reaction coordinate (the critical coordinate in unimolecular
vocabulary), proposes the following three-step process to describe unimolecular decomposition
A+A~A*+A
A*
At ----'" products
In the first step the molecule is collision ally excited. There is then intramolecular energy transfer as A* undergoes various conformational changes.
Among the states of A * is a subset that describes the critical conformation
A t, defined so that there is an even chance that molecules in the configuration At react. (29) The distinction between the isoenergetic A * and At is
apparent in considering the reaction cyclopropane ----'" propene. Unless
sufficient energy resides in a C-C vibrational mode a cyclopropane molecule
does not isomerize, no matter how energized it is. As energy is redistributed
the molecule evolves into the critical configuration from which it may
react or revert back to states of the activated molecule.
The RRKM theory proposes a model to describe energy transfer among
molecules belonging to the classes A* and At. It assumes the following:
(i) In collisional processes that form A * molecules of energy , molecules are randomly distributed among all degenerate states at that energy
(strong collision assumption).
(ii) Vibrational energy redistribution within states of A * is rapid with
respect to the rate of unimolecular reaction (random access assumption).
(29)
(30)
Sec. 9.7
295
Total En ergy
r t .L
T- . .
r
-i r-
v t
v
_
o -
J.
Fig. 9.9. Diagram for energy disposition in RRKM theory.
lifetimes at very low pressure are ~1O-6 sec, much l.)nger than the decomposition
time.
(32) Note that lOr and lOr t are not necessarily equal. Although angular momentum is conserved
as A* evolves to At, the rotational energy may change since the molecular structure is
not invariant. The separation of rotational and vibrational degrees of freedom is only
approximate in both configurations.
296
Chap. 9
where qi is the partition function for the internal degrees of freedom of the
molecule.
To compute k(E) consider the basic RRKM assumptions. They imply
no preference for populating states of the same energy. Thus the states of
A * and At are populated in proportion to their degeneracies. At an energy
E and a specified angular momentum the ratio [At]J[A*] is determined
statistically
[A t] _ g/N(Et)
[A*] - grN(Ev)
where N(Et) is the density of vibrational states of At molecules with vibrational energy Et and likewise N(Ev) for A * molecules with vibrational energy
Ev. The rotational degeneracies gr t and gr are equal since angular momentum
is conserved; as a result no rotational transition can occur as the molecule
evolves from A * to At. The density of states N(Et) is further decomposed
since the critical mode is presumed to be distinguishable from all other
vibrational modes of At,
N(Et) =
L g(E/)N:(E t -
E/ = E/)
.~~o
Here Nc(E/) is the density of states of the critical mode at energy } and
g( Ev t) is the degeneracy of the other vibrational modes at energy Ev t. This
decomposition, treating the critical coordinate as continuous and all other
vibrational modes as discrete, is analogous to distinguishing a specific
reaction coordinate in activated complex theory. Defining an energydependent frequency of passage from A t to products, Y(Ec t), the expression
for the reaction rate constant of A * molecules is
(9.34)
The factor of ! accounts for the fact that half the At molecules revert
to A * molecules. The bracketed expression is the probability that a molecule
with total energy E and vibrational energy Ev has an energy Ec t concentrated
in the critical coordinate.
The product Y(E/)Nc(Ec t) can be estimated from uncertainty principle
arguments. A density of states Nc(E/) implies a separation between levels
--.-1/Nc(E/) and a corresponding energy uncertainty -;5IJNcCE/). Since
(lifetime) X (energy uncertainty) ,....., h
Sec. 9.7
297
we find that(33)
The lifetime is twice the reciprocal of the passage frequency V(E/). (34) Then
_1_
hqrqv
+ k(E)/Y,Bp
(9.37)
Since the sum over vibrational states of At is zero unless Er + Ev = E > EO,
the range of integration has been formally extended. It is generally more
(33)
Careful analysis is required to demonstrate that h, and not Ii, appears here. O. K. Rice,
J. Phys. Chern. 65, 1588 (1961).
(34)
(35)
298
Chap . .9
convenient to express k* in terms of the properties of the critical configuration. Since the rotational degeneracy of both A * and A t are the same we
find
If we assume Eo is the same for all states in the critical configuration then,
incorporating the energy constraints of Fig. 9.9, k* is
which is precisely the result of collision theory since the bracketed expression
is the fraction of molecules with E > Eo. To relate ko to the properties of
the critical configuration is more difficult and can only be done approximately. Various prescriptions have been suggested. (36)
(36)
See W. L. Rase, in Modern Theoretical Chemistry, W. R. Miller, ed. (New York: Plenum Press, 1976), Vol. 2, Part B, Chapter 3.
Sec. 9.8
299
-->.
-->.
(37)
(38)
(39)
(40)
(41)
300
Chap. 9
log,o p (otm)
Fig. 9.10. Pressure dependence of the unimolecular rate constant for CHaNC isomerization.
For clarity the 533 K curve is displaced one log p unit to the left and the 473 K curve is
displaced one log p unit to the right. Solid lines are calculated values based upon RRKM
theory, adjusted to coincide with experiment when k* /koo = 0.1. [Adapted from F. W.
Schneider and B. S. Rabinovitch, J. Arn. Chern. Soc. 84, 4215 (1962).1
d (nm)
473
504
0.43
0.47
533
0.44
D. L. Bunker, and W. L. Hase, J. Chern. Phys. 59, 4621 (1973); H. L. Harris and D. L.
Bunker, Chern. Phys. Lett. 11,433 (1971).
Sec. 9.8
301
there may not be facile energy transfer between states of the critical configuration and some of the states of the activated molecule. How this is
to be interpreted in light of the agreement depicted in Fig. 9.10 is, at present,
unclear. Possibly a quantum treatment of energy flow would establish
CHaNC as obeying the RRKM assumptions. Alternatively, it may be that
the RRKM approach is not sufficiently sensitive to its own assumptions,
at least for the CHaNC isomerization reaction. The fact that the falloff
ratio, for isomerization reaction, is not too dependent on the detailed
structure of the critical configuration(4a) gives this idea some credibility.
If the RRKM assumption of random access is invalid, different models
must be formulated in order to obtain an expression for k(c), if (9.31)
is to be used. Slater(44) investigated the other extreme model and assumed
no vibrational energy transfer between normal modes of a molecule. (45).
This approach is too restrictive. Recent work attempts to steer a middle
(and more difficult) course. (46) The dynamics of vibrational energy transfer
are incorporated in the theory. Effects attributable to vibrational excitation
of modes both strongly and weakly coupled to the critical coordinate can
then be exhibited.
Experimental tests of the RRKM hypotheses have also been developed.
By means of chemical activation molecules are formed with sufficient
energy to undergo unimolecular processes directly. Typical activation
processes are H-atom addition to olefins, radical recombination, and CH 2
addition or insertion. Molecules formed in this way have non-Boltzmann
energy distributions; their rate of decomposition can be used to test the
random access hypothesis introduced previously. A rather different formulation for the rate constant is needed. (47) The decomposition of C 2H 5, formed
by chemical activation, can be accounted for using the same parameters
for the critical configuration that were used to fit the thermal decomposition
(43)
(44)
(45)
(46)
(47)
302
Chap. 9
data. (48) More often, thermal and chemical activation experiments cannot
be carried out on the same molecule. Instead, experiments are performed
on a homologous series for which a single critical configuration is suitable
throughout. The most extensive kinetic studies, involving chemiactivated
radicals formed by H-atom addition to olefins, can be rationalized using the
RRKM approach. (49)
Sec. 9.9
Thermodynamic Analogy
303
The ratio niBjnAnB has the form of an equilibrium constant for formation
of activated complex. For systems in which the activity of the various
species equals their concentration (dilute gases, dilute solutions of nonelectrolytes) (9.41) defines a totally empirical function, the activation equilibrium constant Kt(T),
(9.42)
Continuing in this vein, the corresponding enthalpy, entropy, and Gibbs
free energy of activation are
kT
-h- exp( -!JGotjRT)
kT
-h- exp(!JSotjR) exp( -!JHotjRT)
(9.43)
These quasi-thermodynamic parameters can be related to the Arrhenius
activation energy by means of (8.7); using (9.42) the result is
Ea
dIn Kt
RT + RP --;:::::dT
RT + !JUot
!JHot
+ RT(l
!JHot
+ RT -
- !Jngt)
!J(PV)t
(9.44)
where !Jngt is the change in the number of moles of gas in the activation
process. For gas-phase binary reactions !Jng t = -1.
(50)
Units must be treated carefully in (9.43). The left-hand side has dimensions of a secondorder rate constant while the right-hand side appears to have dimension (-1. A choice
of standard state is implicit in the definition of the free energy of activation, L1 Go t ..
For details see A. C. Norris, J. Chern. Ed. 48, 797 (1971).
Chap. 9
304
iJHot (kJ)
Reference
232
+11.7
220
+16.5
Cyclopropane ~ Propene
270
+50.0
Cyclobutene ~ trans-Butadiene
133
+15.6
24
-20.4
3.3
+26.5
-8.3
+19.4
7.9
-80.1
35.1
-49.7
33.5
-50.1
92.5
-70.9
H + C.H. ~ C.H5
CHa + CHa ~ C2H.
CHa + CO ~ CHaCO
a Standard state for LlSo" for binary reactions is a concentration of 1 mol em-a.
o S. W. Benson and H. E. O'Neal, Kinetic Data on Gas Phase Unimolecular Reactions (Washington: National Bureau of Standards, 1970).
C J. A. Kerr and M. J. Parsonage, Evaluated Kinetic Data on Gas Phase Addition Reactions
(London: Butterworths, 1972).
d V. N. Kondratiev, Rate Constants 0/ Gas Phase Reactions, translation by L. J. Hoitschlag
(Washington: National Bureau of Standards, 1972).
S. W. Benson, The Foundations o/Chemical Kinetics (New York: McGraw-Hill, 1960).
Sec. 9.10
Gas-Phase Reactions
305
/\
The intermediate has greater rotational and vibrational freedom than
either reactant or product. Thus it is not surprising that LlSo+> LlSrxn ;
the values are 50 and 28 J K -1, respectively.
The numerical values of LlSo+ for binary reactions have no particular
significance since they depend critically on the choice of reference state.
It is no longer meaningful to suggest that LlSo+> 0 indicates an activated
complex that is less ordered than the reactants. On the other hand, trends
are important. The more negative the value of LlSo+, the more severe are
the orientational requirements for reaction to occur, i.e., the pass on the
potential-energy surface is narrow. The more positive the value of LlSo+,
the fewer are the orientational constraints corresponding to a much broader
pass on the potential-energy surface. One possible explanation for positive
LlSo+is that, in the activated complex, the reactants are very loosely bound
leading to a large moment of inertia; this is assumed to be the case for
CRa dimerization.
The trends in the other bimolecular examples are reasonably understandable. There are fewer orientational constraints in R-atom addition
to ethylene than in methyl radical addition; the values of LlSo+are in accord
with this interpretation. The hydrogen-transfer reactions of methyl with
ethane and ethylene have the same entropy of activation, indicative of
similar orientational requirements. The addition of CR 3 to CO has a large
negative LI So +, reflecting the severe geometric constraints necessary for this
reaction to be possible.
It is the Arrhenius A-factor which is related to LlSo+and thus to mechanistic models for the structure of the activated complex. The value of
LlHo+ is determined by the barrier to reaction and it cannot be simply
related to a model for the reaction intermediate.
Alternatively, the thermodynamic analogy can be used to provide
estimates of rate constants. In equilibrium thermodynamics one uses
306
Chap. 9
have been studied with X = Cl-, Br-, and NCS-. They are interesting
because, with each ligand, two types of exchange kinetics are found. (52)
Four of the water molecules exchange slowly; iJSot is between 0 and 12
J K-i. The fifth exchanges much more rapidly; iJSot is between 57 and
90 J K -1. These observations may be related to a possible exchange mechanism.
The complex has the octahedral structure shown in Fig. 9.11 a. Exchange
of water may be imagined as occurring through a variety of intermediate
structures; the most extreme examples are shown in Figs. 9.11b and c.
In the first case an extra water molecule is bound before another molecule
leaves; the activated complex is seven coordinate. In the second case a
water molecule is lost to solvent in the activation process; the complex is
five coordinate. The difference between the two intermediates is striking.
Two water molecules bound to the seven-coordinate complex are free to
mingle with the solvent if the complex is five coordinate. The seven-coordinate complex is thus the more ordered and correspondingly should have
a lower value of iJSot. As four of the water molecules exchange with iJSot
< 12 J K-1 we can propose that, for cis exchange, a seven-coordinate
intermediate forms. For trans exchange, on the other hand, a water molecule
leaves before another one is attached; the intermediate is five coordinate.
(51)
(52)
Sec. 9.12
307
(c)
(b)
(a)
Fig. 9.11. Structures of the (a) octahedral CrS 6 X complex ion; (b) seven-coordinate intermediate; (c) five-coordinate intermediate. The solvent S is water.
t
nAB
nAnB
Kt(T) YAYB
yiB
Chap. 9
308
where ZA is the ionic charge and I is the ionic strength of the solution
1= L Zi2Ci. In water at 25C the constant B is 1.0. Since the activated
complex in the reaction
BZB
of ionic strength on the rate constant. From (9.46) and (9.47) we have
log k 2(T)
log koCT) -
tBI1!2[zA2
+ ZB 2 -
(ZA
+ ZB)2]
(9.48)
VII
-0.6~-----L------~----~~
0.1
0.2
If (M1/2)
0.3
Some reactions have not been balanced. [Adapted from S. W. Benson, The Foundations of
Chemical Kinetics (New York: McGraw-Hill, 1960), p. 525, where original references are
given.]
Sec. 9.13
Diffusion-Limited Kinetics
309
in Section 5.7. Some data illustrating the effect of ionic strength are plotted
in Fig. 9.12. The expression (9.48) accounts for the observations at ionic
strengths where Debye-Hiickel theory is reliable.(53)
The effect is readily understood qualitatively. Increasing ionic strength
always reduces the magnitude of the electrostatic effects since the ions
become more shielded by their ion atmospheres. Thus repulsion between
ions of like sign is reduced allowing them to approach one another more
easily; the result is an increase in k2(T). Attraction between ions of opposite
sign is also reduced; they approach one another less readily and k 2(T) falls.
(54)
Ionic strength can affect reactions of neutral species. An example is the reaction of
CO(aq) with hemoglobin [L. H. Parkhurst and Q. H. Gibson, J. Bioi. Chem. 242, 5762
(1967)]. Here changing ionic strength greatly alters the activity of the dissolved CO
thus changing the rate constant.
The treatment in this section is modeled after that given by I. Amdur and G. G.
Hammes, Chemical Kinetics (New York: McGraw-Hill, 1966), pp. 59-64.
Chap. 9
310
(0)
( b)
Fig. 9.13. Model for AB relative diffusion through solvent. (a) A and B separated by solvent
sheath. (b) A and B have diffused into contact when reaction occurs; the contact
distance is d.
{d,
(9.49)
00,
dnA
-D ( - dr
dUA )
+ {lnA-dr
(9.50)
Sec. 9.13
lit
Diffusion-Limited Kinetics
311
[nA
exp({3UA )] = const
(9.52)
(9.53)
where we have used the boundary conditions (9.49) and the fact that the
A-B interaction UA(r) must be zero at large A-B separation. Since h
represents the rate at which A molecules reach the surface of the representative B molecule, it is the rate of reaction of A per B molecule. Multiplying
by nBo to account for the population of B we find the rate law,
where DA
+ DB
from which the rate constant can be found. Expressing it in the conventional units M-l sec- 1 yields the rate constant for formation of an encounter
complex,
(9.55)
with d and the D's measured in SI units.
This treatment, an elaboration by Debye of an approach introduced
by Smoluchowski, (55) is based upon a succession of rather gross assumptions.
The equations which describe bulk diffusion were assumed to remain accurate even at distances where A-B interaction is significant. We then im(55)
Chap. 9
312
r=
00,
nAB =
d,
nAB
IjLl V, LI V
= 4nd j3
3
(9.56)
since at infinite A-B separation the complex does not exist and at contact
there is one encounter complex in a volume "-'4nd 3j3. The arguments which
led to (9.53) can be modified to yield
lAB
(57)
(58)
The first derivation of (9.57) was given by M. Eigen, Z. Phys. Chern. NFl, 176 (1954).
R. M. Fuoss, J. Am. Chern. Soc. 80, 5059 (1958).
R. M. Noyes, Prog. React. Kinetics 1, 129 (1961).
Sec. 9.13
Diffusion-Limited Kinetics
313
reactive). Since molecules remain together for,..."" 10-11 sec there is time for
,..."" 10 rotations, which would seem to allow a sampling of many orientations.
However, if the reactants are fairly strongly associated with the surrounding
solvent, any such rotation would be significantly hindered, and the reaction
rate correspondingly less than the limiting value for spherical molecules.
The most common application of (9.55) and (9.57) is to the problem
of ionic reactions in solution. If the solution is very dilute the AB potential
energy is
(9.59)
where E is the dielectric constant of the bulk solvent. For this simple potential (9.54) can be integrated with the result
(9.60)
The value of kdiff can then be estimated using (9.55). From the NernstEinstein equation (3.13) and the data of Table 3.1 we find that ion diffusion
coefficients at infinite dilution are ,...",,0.6-2 X 10-9 m 2 sec- 1 at 25C (except
for H+ and OH- which are 9.3 X 10-9 and 5.3 X 10-9, respectively). Typical
encounter distances are ,...",,0.5 nm and in water at 25C E"""" 80. Values of
kdiff computed on the basis of these assumptions are plotted in Fig. 9.14
II
10
/
..Q
.....-
--
"'"Q
0>
..... ",,11
o //
o/if
/0/
0
/
/
(/)
::ii!
'"
.4'
J>" "
",-6
Reaction Type
Experimental
Data
Calculated
Values
H+ + B
------
0 A+B
/I
---
o A + OH-
H++OW
-2
-4
ZAZe
Fig. 9.14. Comparison of experimental and calculated values of k dilf for diffusion-limited
reactions. Values calculated using (9.55).
Chap. 9
314
+ B ~ (A
kd!ff
kdlss
k.
B)~products
so that unless k2 ~ k diss the measured value of the apparent rate conStant
does not depend on kdiff alone. While k2 is hard to estimate, the related
quantity k' = k2kdifr/kdiss can be given a physical interpretation. It is the
intrinsic rate constant for the A + B reaction, i.e., the rate constant which
would be observed if the reactants were placed in contact. (60) As such k'
(59)
(60)
It is worth noting that the method of calculation must be significantly altered for
more concentrated solutions to insure that the values of k dlfr found by letting A diffuse
toward a stationary B or vice versa be the same. See L. Bass and W. J. Greenhalgh,
Trans. Faraday Soc. 62, 715 (1966); A. M. Watts, Trans. Faraday Soc. 62, 2219, 3189
(1966).
R. M. Noyes, Prog. React. Kinetics 1,129 (1961); T. R. Waite, J. Chern. Phys. 28,103
(1958); 32, 21 (1960).
Sec. 9.13
Diffusion-Limited Kinetics
315
In general, given the uncertainty in the estimate of kdiff' it is hard to disentangle chemical and diffusive contributions to k app . (61)
A steady state is not established instantaneously. After A and B form
an encounter complex and react, the mean concentrations of unreacted
A and B in the vicinity are below their steady state values. Diffusion of
A and B is required to reestablish the steady state. As a result there is an
induction period before steady-state behavior is observed. At shorter times
the binary rate constant, k app , is time dependent, and greater than its
steady-state value, (62)
kapp(t) =
k'
k'kdiff
{
kdiff
1
+ 1 + kdifrlk'
(9.62)
Choosing the same values for d and D that were used in estimating kdiff
we find that, as long as the reaction is diffusion limited (k' > k diff), ",-,10- 7
sec is required for kapp to differ by less than I %from its steady-state value.
After "'-' 10-9 sec the difference is "'-' 10%.
The effect is significant for reactions with short half-times. In an
ordinary diffusion-limited reaction under conditions where the species
B is in great excess, the half-time is Ij(kdifdBD. If k diff "" 1010 M-1 sec-I,
a steady-state treatment of the kinetics will suffice as long as [B] ;:S 0.1 M. (63)
If [B] is larger, the effect of transient behavior cannot be ignored. Another
instance where transients are important is in the kinetics of fluorescence
quenching. Here, the intrinsic lifetimes of the photo excited species are
typically "'-' 10-9 sec. The photostationary state is not established and
effects attributable to the time dependence of kapp can be observed.
For the practicing kineticist this is usually not an issue since the major significance of
the diffusion limit is that derived binary rate constants cannot exceed kdl!!. The
individual contributions to kapp have been distinguished in an analysis of the kinetics
of iodine-atom recombination in CCI, [R. M. Noyes, J. Am. Chern. Soc. 86, 4529
(1964)].
(62) R. M. Noyes, Prog. React. Kinetics 1,129 (1961); T. R. Waite, J. Chern. Phys. 28,103
(1958); 32, 21 (1960).
(63) The measured rate constants will be "'" 10% too large, but efforts to obtain greater accuracy are usually unwarranted.
(61)
Chap. 9
316
~ exp ( - :~)
(9.AI)
where En are the energy levels of the molecule of interest. The thermodynamic consequence is that q(T) determines the molar chemical potential
via the relation
(9.A2)
fl = -RTln q(T)/N
Thus, if a chemical equilibrium is established in the reaction
A+
we know that
flA
B~AB
+ flB =
flAB
'LIn =
n
A fuller discussion can be found in texts on this subject, e.g., T. L. Hill, Introduction to
Statistical Thermodynamics (Reading, Mass.: Addison-Wesley, 1960), Chapters 4,
8, and 9.
Appendix A
317
The consequence is that the partition function (9.Al) may be factored into
a product of contributions from each degree of freedom
(9.A5)
where each q is now calculated by an expression like (9.AI); the energy
levels now refer to those corresponding to a particular degree of freedom.
When expressions for the energy levels, deduced from quantum mechanics, are introduced into (9.Al) the following results are obtained:
_ ( 2nmkT )3/2
qt h2
V
qr
Sn 2 IkT
ah 2
Sn2(ABC)1!2(2nkT)3/2
(linear molecules)
(nonlinear molecules)
ah 3
=
v
qe
11
i
exp( -hvi/2kT)
[1 - exp( -hVi/kT)]
II [2 .smh (hVi)]-l
2kT
rigid
rotors
harmonic
oscillators
(9.A6)
Chap. 9
318
drp dx
f --he-</kT
(9.A7)
for one classical degree of freedom. An integration over position (x) and
momentum (p) replaces the sum over energy levels. The energy 10 is given
by the classical expression
10 =
f~ + Vex)
the sum of kinetic and potential energies. The factor h is introduced so that
classical and quantum mechanics correspond in the limit where they are
known to lead to the same predictions; if it were not included the classical
partition function (9.A7) would also not be dimensionless.
(9.BI)
(9.B2)
(9.B3)
q/
<t
r
o <10/ <
101 t
where g / and
tit
<
Ev t
Elt
<
102 t,
etc.
Problems
319
L g/ roo
J ,.t
'i
exp( -(JEt) dE t
7f ~ g/ exp( -(JE/) =
(9.B4)
kTq/
'i
Problems
9.1.
~4()()
CI
+ H. ~ HCI + H
is 1.2 X 10 13 cm 3 mole-'sec-'. Calculate the A-factor for this reaction assuming a linear transition state with bonds 300/0 longer than in H2 or HC!.
Assume that the vibrational frequencies in the activated complex are sufficiently large so that none are excited. The bond lengths in H2 and HCI are
0.0742 nm and 0.1275 nm, respectively. The electronic degeneracies are gel
= 6, gH = 1, gt = 2. (Estimation of the vibrational frequencies of HHCl
is not required.)
9.3.
For the reaction of Problem 9.2 compute the A-factor assuming the CIHH
angle to be 90. For an XYZ molecule with such geometry the moment of
inertia product is
+ I")(lllI,,
ABC
(Ill
III
mx(my
122
where
- I;,)
+ mz)R,'/ M
mz(mx + my)R.'/M
I., = mxmzR,R./ M
Chap. 9
320
Does the Eyring theory allow discrimination between the two extreme
geometries for the CI + H. reaction?
9.4.
Recalculate the A-factors in Problems 9.2 and 9.3 assuming no bond expansion in the activated complex. Does this alter your choice for the more
probable geometry?
(a) Using the data of Table 9.4 determine n+ in the isobutene + HCI
reaction at 400 K; the pressure of each of the reactants is 1 atm. To what
molecular concentration does your value of n+ correspond?
(b) How do you define the activated complex? Interpret the result of part
(a) within this context.
9.8.
LJSo+is much larger for the reaction cyclopropane -' propene than for the
reaction cyclobutene -' trans-butadiene. Why?
9.9.
Look up So for HCI(g), C.H,(g), and C.H 5CI(g) at 298 K. Use these values
and the data of Table 9.4 to compute LJSo+for the reaction
Compare your value with that for the addition of HCI to isobutene. Comment on the difference.
9.10. (a) Extend the thermodynamic analogy of Section 9.9 to define an activation
volume LJ Vo+ in terms of the pressure dependence of the rate constant.
(b) The pressure dependence of the rate constant for the solvent exchange
reaction
Problems
321
3.82
4.01
4.56
4.95
5.79,5.76
6.00
0.001
0.059
1.027
2.02
3.31
4.00
What is Ll Vol?
(c) Would you expect any correlation between Ll Vol and LlSot? Explain.
9.11. Stokes' law permits an estimate to be made of ionic radii. Combine this
with the Nernst-Einstein relation (3.13) and the expression for the diffusion
limited rate constant to show that
and that
Use these results with the data of Table 3.1 to compute kdtrr and d for the
reaction, at 25C,
H+ + OH- ~H20
At this temperature 'YJ = 0.89 X IO- a kg m- I sec-I and
mental value of kdtrr is 1.4 X 1011 M -I sec-I.
E =
9.12. Carrv out the same calculation described in Problem 9.11 to estimate d
and kdtrr for the reactions
+ OH- ~ NH,OH
+ HCO a- ~ COa 2- + H,O
+ HCO a- ~CO, + H 20
NH,+
OHH+
Compare your results with the experimental data quoted in Table 5.3.
9.13. The general theory of diffusion-limited rate constants may be combined
with the theory of the primary salt effect to estimate the effect of ionic
strength on k dW Assume mean diffusion coefficients 1.3 x 10- 9 m' sec I
and encounter distances of 0.5 nm for ions in water at 25C.
(a) Use (9.55), (9.59), and (9.60) to show that, when I
0,
(b) Combine this with (9.48) and estimate k,(T) for the reaction
A ZA
+B
ZB
~ product
ZA
= 1,2,3,
ZB
= 1,
Chap. 9
322
(c) From the formula of (b), determine the value of I at which kdlt! is
the same if the ionic charge product ZAZB is +q or -q.
(d) What does the result of (c) suggest about the range of validity of the
general formula deduced in (b)?
9.14. The RRKM model may be treated phenomenologically as
<u)
2kT
v+
CI
CI
k,/k.,
+ H2 ~ HCI + H
+ T2 ~TCI + T
Determine the isotope effect for the reverse reactions. The vibrational
frequencies are 1.32 x 1014 sec- 1 for H. and 8.97 x 1013 sec- 1 for HCI. Those
for T. and TCI can be estimated since
and the force constant is determined by the shape of the potential surface.
9.17. Demonstrate the equation for
k d188 ,
(9.57).
9.18. Estimate kd188 as a function of ZAZB for ionic complexes in water at 25C.
Assume D A + DB = 2.6 X 10-9 mO sec- 1 and d = 0.5 nm.
9.19. What is
Keq
+ Ac- =
General References
323
Assume d = 0.4 nm, 0.5 nm, and 0.6 nm. At what separation is Keq (a)
maximum, (b) minimum? Can you ascribe physical significance to these
results?
9.20. Singlet-singlet excitation transfer, as discussed in Section 6.5.2, takes place
at distances of 6-7 nm. Use data from Problem 9.13 and estimate the
diffusion limit for this process in water. (The largest values observed are
~ 3 x 10 11 M -1 sec- 1 .) Is the process diffusion-limited?
9.21. (a) Use data from Problems 9.13 and 9.20 to estimate the time required to
establish a steady state when studying singlet-singlet excitation transfer.
(b) Would transient behavior complicate studies made using pulsed-laser
flash photolysis?
9.22. As the intrinsic rate constant varies, the time required to establish a steady
state does also. For reactions for which the measured rate constant is less
than ! the diffusion limit, under what conditions, if any, does non-steadystate behavior complicate the interpretation of an experiment?
General References
I. Amdur and G. G. Hammes, Chemical Kinetics (New York: McGraw-Hill, 1966), Chap-
ters 2 and 4.
H. S. Johnston, Gas Phase Reaction Rate Theory (New York: Ronald Press, 1966).
K. J. Laidler, Theories of Chemical Reaction Rates (New York: McGraw-Hill, 1969),
Chapters 2, 3, and 6.
Potential-Energy Surfaces
D. G. Truhlar and R. E. Wyatt, Adv. Chem. Phys. 36, 141 (1977).
10
A Dynamical Model
for Chemical Reaction
10.1. Introduction
If the reaction cross section, Q(E, i, j), is known, the rate constant for
the corresponding chemical reaction can be calculated from (8.6) and (8.11).
To interpret the details of molecular beam experiments even more information is needed. The hard-sphere model was obviously far too naive to give
reliable estimates of the reaction cross section. Improvement, by considering
the dynamics of reaction across a realistic potential-energy surface such as
that described in Section 9.2, is a formidable quantum mechanical problem.
As already mentioned it has not been solved, except for low-energy H + H2
chemical reaction.
In this chapter we shall try to clarify this chemically significant problem
by studying the dynamics of the collision process A + BC in extremely
simplified terms'!!) We shall not attempt to calculate Q(E, i, j). Instead we
shall see how changing various molecular parameters influences reaction
probability and affects energy transfer. The effect of varying the collision
energy, the reactant mass, and the shape of the reaction profile will be
investigated.
The results of our analysis will show that slight variations in the nature
of the ABC potential-energy surface, in the collision energy of the reactants,
or in the initial vibrational energy of the BC molecule can have dramatic
effects on the collision dynamics. In our model we shall emphasize collinear
(1)
The treatment is adapted from B. H. Mahan, J. Chern. Ed. 51, 308, 377 (1974).
325
326
Chap. 10
collisions and ignore effects due to rotational motion. The model, while
highly restrictive, nonetheless illustrates many aspects of reaction dynamics.
The fact that so many features of chemical reaction can be reproduced
by a simple model emphasizes the point that chemical reaction is not a
mysterious problem of rearrangement and recombination. It is, on the
other hand, directly determined by the type of collision taking place and
the nature of the interparticle forces.
Sec. 10.2
327
X -+-----01
collisions are possible. Many features of reaction dynamics can be illustrated even if the model is simplified further and A, B, and C are assumed
to obey the laws of classical mechanics.
For the collinear system the total kinetic energy of relative motion is
the sum of the vibrational kinetic energy of the BC diatom and the kinetic
energy of A relative to the center of mass of BC Letting mA, mB, and mc
denote the atomic masses yields
(10. I)
Here mBmC/(mB -+- me) is the reduced mass for BC vibration and
mA(mB -+- me)/(mA -+- mB -+- me) the reduced mass for A-BC relative
(10.2)
X - yY
x = X and y = Y/n
with n = [mA(mB -+- mc)2/mBmcM]1/2, where M
kinetic energy becomes
(10.3)
= mA
-+-
mE
Chap. 10
328
......-
rBe
= constant
x
Fig. 10.3. Mapping of molecular coordinate system into Cartesian (x, y) system. Lines
of constant r AB and rBC intersect at an angle f3 = tan- 1(m BM/mAmc)1/2.
rAB
(10.5)
Lines of constant rBe run parallel to the x axis and lines of constant r AB
have a slope of (Y1'))-l in the x-y coordinate system; the angle between
these lines is determined by the atomic masses
(10.6)
The mapping is shown in Fig. 10.3.
Sec. 10.3
329
00
rAB
V=
+ 00
rr-2t!
V=D
refraction
/ //(diatom dissociates)
~
step
v=o
f3
v=+oo
reflection
(diatom remains bound)
wall 2
Fig. 10.5. Potential-energy surface and trajectory of representative particle for collinear
hard-sphere A interacting with square-well oscillator Be.
Chap. 10
330
= v cos(n -
2~)=
-v cos
2~
where v is the magnitude of the initial velocity vector. Using (10.3) and
(10.1), the kinetic energy of relative A-BC motion after collision is
E'
mA(mB
+ me)
2M
v 2 cos 2
2~
Eo cos 2 2~
where Eo is the initial kinetic energy of the ABC system, which, in this
example, is the initial relative A-BC kinetic energy. If BC remains bound
the rest of the energy resides in the vibrational mode of BC, Ev':
Ev' = LIE = Eo - E' = Eo sin 2
2~
4mAmO
+ me)
--~-----:---'---:-
mB(mB
If; ~
4mBme
(mB
me)2 '
The mass effect may be roughly correlated with the activation (deactivation)
efficiency of buffer gases in unimolecular decompositions (bimolecular
recombinations ).
As an example, catalysis efficiency in the isomerization of cyclopropane
to propene can be interpreted, in part, in terms of the mass of the buffer-gas
molecules. To activate CaH6' energy is most likely transferred first to a
C-H bond by collision of a buffer-gas molecule with the H atom; the extra
Sec. 10.3
331
Efficiency computed
from (10.7)
Relative efficiency in
cyclo-C 3 H. isomerization
from Table 5.1
4 (He)
0.77
0.060
40 (Ar)
0.35
0.053
2 (H 2 )
0.95
0.24
28 (N 2 )
0.39
0.060
energy can then be partitioned among other vibrational modes and eventually weaken a C-C bond. From (10.7) the relative energy transfer may be
estimated by considering the head-on collision of a buffer gas A with the
CH moiety so that B = I and C = 12. The ratio iJEj Eo, computed for
buffer gases of mass 4, 40, 2, and 28 (He, Ar, H 2 , N 2 ), is tabulated in
Table 10.1; the results are in qualitative agreement with the experimental
activation efficiencies of Table 5.1. A buffer of mass 4 is more efficient
than one of mass 40; one of mass 2 is better than one of mass 28. Considering the approximate nature of the model being used-effects due to
multiple collisions, molecular structure, molecular rotation, and intermolecular forces having all been ignored-no better agreement could be
expected.
The dynamics of the hard-core, square-well problem is particularly
simple for the initial conditions just considered. Sometimes multiple collisions occur as shown in Fig. 1O.6a. The representative particle collides
with wall I more than once, i.e., A hits B, then B hits C, and finally Band
A collide again before atom and diatom separate. If, as in Fig. 1O.6a,
BC is initially vibrationless, multiple collisions can only occur when f3 < 60;
as mB decreases these become more common and are predominant if
f3 < 50. Such repeated collisions significantly increase the energy transfer;
instead of (10.7) the inelasticity is
iJE
Eo
sin 2 2nf3
(10.8)
332
Chap. 10
Fig. 10.6. Possible trajectories with potential-energy surface of Fig. 10.5. (a) Multiple AB
encounter. (b) Be has initial vibrational energy. (c) Be has initial vibrational energy and
multiple AB collisions occur. (d) Be has zero final vibrational energy.
axis. Two such trajectories are shown in Fig. 1O.6b and c. The energytransfer calculation is more complicated but, on the average, less energy
is transferred to the vibrational mode. The possibility of multiple collisions
is much greater. For some trajectories, such as the one shown in Fig. 1O.6d,
collision leads to vibrational deactivation, which affects bimolecular recombination kinetics.
In bimolecular recombinations the vibrational energy is much larger
than the relative translational energy. Under such circumstances (10.8)
also describes the fractional energy transfer from vibration to translation
(see Problem 10.2). Let us consider the reaction
Br
+ Br + M ~ Br2 + M
Sec. lOA
333
If the interaction were repulsive (.1 > 0), the inelasticity would decrease;
if E t , the relative A-Be translational energy, is less than .1, the particle
cannot overcome the barrier and is reflected at the discontinuity.
If A is preferentially attracted to B, assume the attraction is to B alone
and modify the potential surface by introducing a step at constant r AB
as shown in Fig. 1O.7b. For the initial conditions shown the particle is
refracted upon entering the well, reflected at wall I, and refracted again
upon leaving the well. Geometric considerations show that the well has
not altered the energy transfer. The inelasticity is still given by (10.7) and
(10.8).
The other limiting case occurs when A is attracted to e, which can
be modeled by incorporating a potential well with a discontinuity at a
constant separation r AC as shown in Fig. 1O.7c. The inelasticity calculations
(3)
Chap. 10
334
AE
= (E- il)sin2
2/3
AE
= E sin 2
AE
211
=0
Fig. 10.7. Potential-energy surface features which complicate trajectory analysis; .1 is the
depth of the potential well. (a) Inelasticity altered due to potential step perpendicular to
x axis. (b) If A only attracted to B, .1 Ev is not affected. (c) A is attracted to C. (d) If potential
well is deep enough temporary trapping occurs.
for the trajectory shown are no longer simple. There is, however, an increase
in energy transfer when the interaction is attractive.
The atom-diatom attraction not only affects energy transfer but it
also introduces the possibility of "sticky" collisions. A possible trajectory
is shown in Fig. 1O.7d. After reflection at wall 1 the representative particle
attempts to recross the step but its x component of kinetic energy is not
sufficient. Here the kinetic energy of the particle is E - L1 while it is in the
well; the condition for crossing the step on the first try is
(E - L1) cos 2 2f3 > I L1 I
(10.10)
Sec. 10.5
335
+ Be ---" AB + C
v =0
Fig. 10.8. A square-trough potential-energy surface for the reaction
A +BC~AB + c.
V=OD
x
A+BC
336
Chap. 10
Fig. 10.9. A portion of the potential-energy surface for the exchange reaction H + H2 ~
H2 + H assuming collinear geometry. The energies are in units of kJ mol-I. [Adapted
from R. E. Weston, Jr., J. Chern. Phys. 31, 892 (1959).]
y/x
(10.11)
where E is the total kinetic energy of the system; E t and Ev are the initial
relative A-BC kinetic energy and BC vibrational energy, respectively.
As long as f3 > 60 there are only three types of reactive trajectories, those
shown in Figs. 1O.lOc,d, and e. Since both y and x are negative at the
reference line in Fig. 1O.lOc, () is positive. In Fig. 10.1 Od () is negative,
since y > 0 and x < 0 at the reference line. In Fig. lO.lOe () is positive for
the two trajectories shown.
Sec. 10.5
(c)
f3
337
tl'=tl- Iel
cptf3tf3' = IT
f3' = f3 tiel
+f3 +f3' = TT
e' t cp = f3
e' t cp = tl
reference line
e ttl = e'
Fig. 10.10. Effects of initial vibrational energy and phase on product formation and energy
distribution. The diagrams at the top show trajectories with no initial vibrational energy
for (a) f3 = 60, (b) f3 > 60 surface. (c) Construction for deducing 8' for collision sequence
where A hits B, then B hits C. (d) Effect of phase on reactivity. (e) Construction for finding
8 when collision sequence is B hits C, then A hits B. (f) Trajectories with 8 = 30 on f3 = 60
surface.
338
Chap. 10
Extent of reaction
degrees
0
0-30
30
30-60
60-90
1 (J' I,
Total
Partial
None
Partial
Total
degrees
1 (J 1
1 (J 1
120 120 -
1 (J 1
1 (J 1
{J'
= {J - 1() I,
rp
+ {J + {J' =
n,
rp+O'={J
(10.12)
and hence
()' =
3{J -
1 () 1 -
(0
(10.13)
The case -n/2 + {J < () < 0 (Fig. 1O.10d) differs only in that {J'
{J + 1 () 1 and thus
()' =
3{J
1 () 1 -
(-n/2
+ (J < () < 0)
(10.14)
n - {J -
1 () 1
+ (J)
(10.15)
Sec. 10.6
Exoergic Reactions
339
E cos 2 ()'
Ll,
E,,'
E sin 2 ()'
(10.16)
Chap. 10
340
v=o
v=o
v=o
v=o
v
Fig. 10.11. Effects of exoergicity (Ll < 0) on product formation and energy distribution.
(a) Step in exit channel corresponding to repulsion between C and center of mass of AB.
Final translational energy altered by step. (b) Step in exit channel due to BC repulsion. Both
translational and vibrational energy of products are affected. (c) Step in entrance channel
due to attraction between A and center of mass of BC. Large y trajectory which would have
led to inelastic collision becomes reactive if E is small. (d) Step in entrance channel. Collision between A and vibrationally excited BC leads to inelastic energy transfer instead of
reaction because of the exoergicity.
the reagent channel; if the reaction were thermo neutral (,1 = 0) the trajectory would be nonreactive. From Fig. 10.5 we see that 'JI = :rr - 2fJ. Since
the total kinetic energy in the region to the left of the step is E - ,1, where
,1 < 0, the x component of the kinetic energy is
(E - ,1) cos 2
'JI
Thus if
(E - ,1) cos 2 2fJ
<
-,1
or E
<
(10.17)
Sec. 10.6
Exoergic Reactions
341
= {
large y
small y
(10.18)
Chap. 10
342
and
sin ()1
(10.19)
The further course of reaction is determined by the angle ()1' Any trajectory
(0)
v=o
reflection
if Et ( Ll
Fig. 10.12. Effect of endoergicity (.d > 0) on product formation and energy distribution.
(a) Step in entrance channel. Reaction only occurs if E t = E cos 2 0 >.d. (b) Step in entrance
channel. Effect of endoergicity in modifying trajectories to decrease reactivity. (c) Step in
product channel. Reaction only occurs if E t ' = E cos 2 0' > .1.
Sec. 10.7
Endoergic Reactions
343
which leads back to the step is necessarily nonreactive. Since 1 ()l 1 > 1 () I,
initial trajectories corresponding to small vibrational energy (small 1() I)
may, after refraction, be redirected along a nonreactive path as can be
seen in Fig. 10. 12b. Thus not only does the barrier prevent reaction if
E t < LI, it also reduces the reaction probability (when Ev is small) for
many trajectories if E t > LI.
A barrier in the product channel (Fig. 10.12c) can only be surmounted
if Ee' = E cos 2 ()' > LI. In this example the endoergicity is due to attraction
between C and the center of mass of AB. Here a small amount of vibrational
energy is far more effective in promoting reaction than is increasing the
translational energy. As long as fJ - nl2 < () < 2fJ - n12, (10.13) or
(10.14) are valid for all fJ and
()' = 3fJ - () - n
so that
Ee'
E cos2(3fJ - ())
E(cos
(10.20)
As usual Ee' depends on the sign of (). If fJ > n13, Ee' is larger for () positive.
Using (10.11) and choosing the phase which maximizes E/, (10.20) becomes
E/
(10.21)
Et
+ H 2+--->.HeH+ + H,
LIE = 77 kJ mol-1
Chap. 10
344
o
2
96
192
288
385
0.06
0.10
0.13
0.17
0.49
0.35
0.31
0.25
1.95
0.93
0.55
0.34
1.70
0.99
0.56
2.35
1.22
0.68
2.49
1.70
0.89
a W. A. Chupka, in Ion Molecule Reactions, J. L. Franklin, ed. (New York: Plenum Press, 1972),
Vol. 1, p. 73.
Sec. 10.8
Noncollinear Geometries
345
greatly complicates the dynamics. (7) Since even for the simpler examples
slight changes in initial conditions may alter the outcome of interaction,
it is clear how difficult it is to deduce a potential-energy surface from the
results of molecular beam experiments.
For the much simpler problem of an intermolecular potential between
argon atoms, discussed in Section 2.8, we saw that only by combining data
from different types of experiments could a reliable potential be constructed. In that case the potential was a function of one variable only,
the argon-argon distance. For the atom-molecule system the potentialenergy surface is a function of the three independently variable atomic
distances. To vary the parameters of such a potential in order to reproduce
molecular beam scattering data is a herculean task; some efforts toward
this goal are discussed in Section 10.9. To carry out a similar program for
reactions involving more complicated molecules is presently beyond consideration. At present reactive scattering data are used to establish the
general features of a potential-energy surface, not the precise details.
(8)
346
Chap. 10
0-
~B
(1)
(i)
'ej .....
(i i)
(iii)
e:e
(II)
.e
(j)
%
-e
(iv)
A"B
(ii)
(ii i)
::s
::s
Sec. 10.8
Noncollinear Geometries
347
3r-------------------------------~
o
2
0 0
100
200
300
+ HD system
If the impact parameter is larger than the hard-core diameter the model is no longer
remotely applicable. The arguments presented are useful at smaller impact parameters.
Chap. 10
348
As a consequence the amount of KrD+ formed in recoil processes is enhanced; the amount of KrH+ is decreased.
The qualitative features of the high-energy cross-section ratio and
scattering patterns can now be understood. Smaller impact parameters
favor formation of KrD+ because of rotation and the requirement of multiple collision for formation of KrH+. The major recoil product is KrD+,
as reflected in its scattering pattern. Recoil processes produce little KrH +;
only large impact parameter processes, which lead to forward scatter,
are effective. The data reflect these considerations. The enhanced cross
section for KrD+ is also understandable. As collision energy increases
large impact parameters are less likely to lead to reaction (Sections 8.5-8.9).
Thus KrH+ formation becomes increasingly improbable. Smaller impact
parameters, which are now more significant, favor KrD+ formation. The
result is an enhancement of Q(KrD+), as observed.
Problem 10.3 illustrates that for collinear trajectories in which the
diatom has no initial vibrational energy (0 = 0), multiple collision is
mandatory if fJ < nj4. According to (10.6) such values of fJ are possible
only if mB < me. At nonzero values of the impact parameter b, since
reaction and separation cannot occur, the Be diatom starts to rotate after
collision of A with the B end of the molecule. The larger the collision energy
or the larger the value of b,cl1l the greater the angular velocity. The second
collision is more likely to involve the heavy end of the molecule. Sufficient
energy transfer is now possible; separation may occur without further
collisions. In general, therefore, the collinear approximation overestimates
effects due to multiple collisions. At nonzero impact parameters more than
two collisions are quite unlikely unless the potential surface favors complex
formation.
(11)
->.
There are limits on the range of b for which these considerations apply. See footnote
(10) in this chapter.
Sec. 10.9
en
f-
349
15
Trajectory Calculation
:l
>-
Experimental Results
~10
a:
!:::
co
a:
~
.J
t9
en
0
J:
-20
20
40
60
strate the accuracy with which scattering patterns can be calculated. (12)
In particular, we shall find that good results can sometimes be obtained
using most approximate model potentials.
(13)
(14)
Another interesting system for which scattering patterns have been calculated and compared with experiment is the reaction of H+ with D 2. Depending upon energy, various
channels are open. The possibilities are: D+ + HD (E> 4 kJ mol-I); H + D2+'
D + HD+ (E> 180 kJ mol-I); H + D+ + D, H+ + D + D (E> 440 kJ mol-I)
[J. R. Krenos, R. K. Preston, R. Wolfgang, and J. C. Tully, J. Chern. Phys. 60, 1634
(1974)].
P. Brumer and M. Karplus, J. Chern. Phys. 54, 4955 (1971).
R. N. Porter and M. Karplus, J. Chern. Phys. 40, 1105 (1964).
350
Chap. 10
EXPERIMENT
KINEMATIC
N Dire<: tlon
D~
MODEL
Direction
Fig. 10.16. Comparison of calculated and observed angle-velocity flux distribution for the
reaction Ar+ + D. ~ ArD+ + D at a collision energy of 260 kJ mol-I. Calculated values
from T. F. George and R. J. Suplinskas, J. Chern. Phys. 54,1037 (1971); experimental data
from M. Chiang, E. A. Gislason, B. H. Mahan, C. W. Tsao, and A. S. Werner, J. Chern.
Phys;52, 2698 (1970). [From Z. Herman and R. Wolfgang, in lon-Molecule Reactions, J.
L. Franklin, ed. (New York: Plenum Press, 1972), Vol. 2, p. 580, reprinted by permission.]
energy surface since the collision energies are only somewhat above threshold.
Problems
351
Problems
Problems 10.1-10.4 are designed to exhibit some general features of nonreactive
collisions assuming the potential-energy surface of Fig. 10.5.
10.1.
Show that for fJ = n/2n all collinear collisions are elastic (no interconversion of translational and vibrational energy).
10.2.
(j
Ev = Eo sin2 (j',
(j'
= n
+ (j -
2mfJ
(c) Compute iJEv/Eo in Br2 recombination when the buffer gas is He,Ne,Ar,
Kr. (The ratio of the recombination rate constants is 1 : 1.4 : 2.2 : 2.9
at ......,300 K.)
(16)
352
10.3.
Chap. 10
(a) Show that the only possible values of f) are -n12 < f) < f3.
(b) For f3 > nl4 show that only one AB collision is possible if f) < 3f3 - n.
(c) For f3 > n/4 show that there are two AB collisions if 2f3 - nl2 < f) < f3.
(d) Energy disposition is determined by I f) I, not f). For I f) I < f3 show
that the fraction of collisions with f) = -I f) I is HI + tan I f) I cot f3].
(e) For 3f3 - n < f) < 2f3 - nl2 there can be 1 or 2 AB collisions. Show
that, of the trajectories with f) in this range, a fraction 2/[1- cot f3 tan(f3 -f))]
leads to only a single collision.
10.6.
(a) For the potential surface of Fig. 1O.l1b calculate E t ' for the trajectory
shown as a function of Eo, Ll, and f3.
(b) Construct a potential-energy surface for which the exothermicity is
due to A-B attraction and determine E t ' for a trajectory where f) = 0
(Ev = 0).
10.7. Assume I Ll I?> Eo and determine Et'/l Ll I and Ev'/l Ll I for the two "repulsive" surfaces of Figs. 1O.lla and b; for the two "attractive" surfaces
of Fig. lO.11c and the one constructed in Problem 10.6b. Note the qualitative differences.
10.8.
353
General References
10.9.
10.10. Consider the surface of Fig. 10.10 with an initial trajectory where () = O.
Modify the analysis used in Problem 10.2 and show that:
(i) all paths are reactive if n/(2n + 1) 2': f3 2': n/(2n
(ii) no paths are reactive if f3 = n/(2n), n 2': 1.
+ 1.5),
n 2': 1;
General References
P. J. Kuntz, in Modern Theoretical Chemistry, W. H. Miller, ed. (New York: Plenum Press,
1976), Vol. 2, Part B, Chapter 2.
R. D. Levine and R. B. Bernstein, Molecular Reaction Dynamics (Oxford: Clarendon Press,
1974), Chapter 6.
B. H. Mahan, Ace. Chern. Res. 8, 55 (1975).
R. N. Porter, Ann. Rev. Phys. Chern. 25, 317 (1974).
Physical Constants
Quantity
Symbol
Value
Avogadro's number
No
6.02252 X 10 23 mol- 1
Boltzmann's constant
1.38054 X 10- 23 J
Electron charge
1.60210 X 10-19 C
Electron mass
me
9.1091 X 10-31 kg
Faraday's number
9.64868 X 104
C mol-1
Gas constant
8.3143
Planck's constant
Velocity of light
Co
355
K-1
K-1
mol-1
Index
Where new material or new concepts are introduced in problems, the page numbers are italic.
Absorption of radiation
coefficient of, 166
quantum mechanical expression for,
167
measurement of intensity, 166
nonstationary states, 173
Acridine fluorescence lifetime, 178
Actinometry, 166-167
Activated complex, 273
and bond expansion, 283
and critical configuration, 294
and pass energy, 273
structure estimation, 284
structure in HI" 283-285
Activated complex theory
critique of assumptions, 281
and depletion, 278
and kinetic isotope effect, 287-291
and local equilibrium, 277
model calculations of rate constants
discrimination between reaction
geometries, 285, 287
hard-sphere model, 282-283
identification of reaction coordinate,
282
in reaction of Hand 12, 283-285
tabulation, 286
and separability of degrees of freedom,
277
and three-body collisions, 278
Activated molecules
and buffer gases, 126
and collision energy, 125
in unimolecular reaction, 124, 294
357
Index
358
Arrhenius activation energy (cont.)
negative values
in radical recombination reaction,
130
Catalysis
by enzymes, 145-147
competitive inhibition, 147
noncompetitive inhibition, 147, 161
uncompetitive inhibition, 147
heterogeneous, 147-152
and adsorption, 148
analogy to enzyme catalysis, 148-149
and chemisorption, 151
and diffusion, 147-148
homogeneous, 142-144
simple mechanism, 144
Catalysts, 142-144
poisoning, 143
and surface chemistry, 151-152
and surface reconstruction, 152
CCIF3, laser-induced decomposition, 192,
198
Index
Chemisorption (cont.)
energetics of, 150
site specificity, 151-152
surface reconstruction, 152
surface specificity, 151
CH3I
beam reaction with K, 248-249
beam reaction with Na, 265
l-Chloroanthracene, excitation transfer to
perylene, 198
CH3NC, isomerization and RRKM theory,
299-302
Clep), reaction with CH4, 99-100, 109
Ch
beam reaction with 2H, 252, 257-258
angle-velocity flux distribution, 258
angular distribution of products, 257
collision geometry of, 258
reaction with H 2, 133
CIO-, reaction with r, 77, 160
CO
laser, 189, 195
reaction with 02, 184, 208
C02 laser, 188, 195
Collision complexes
angular distribution of products, 259
lifetime, 259
Collision dynamics (see also Elastic,
Inelastic, Reactive collision
dynamics)
with collinear geometries, 325-345
with noncollinear geometries, 345-348
Collision energy
and centrifugal barrier, 246-247
and reactivity in beam reactions, 246254
and thermal reaction, 271
Collision frequency, 20-24
as function of relative speed, 235
and mean free path, 20
total, 24
Collision probability, 18-19
Competition methods, 84 (see also Quenching, photochemical)
in kinetics of cyclohexane rearrangement,
86-87
limitations of, 178-179
Concentration waves, 220
Concerted rotation, and orbital symmetry
correlations, 186-187
[Co(NH3)s]202 4+, reaction with V(II), 160
Consecutive reactions, arbitrary order,
120-124
359
Consecutive reactions, arbitrary order
(cont.)
360
Differential cross section, 239
in beam reaction of Na and Br2, 265
Diffusion (see also Transport equations)
and branching-chain reactions, 206
as competition method, 84
and electrocapillarity, 222
and heterogeneous catalysis, 147-148
and oscillating reactions, 220
and periodic precipitation, 221
Diffusion coefficient, 24
of dilute hard-sphere gases, 32-38
of electrolytes, 56
and equivalent conductance, 53-55
and Walden's rule, 35, 59-60
Diffusion-limited rate constant
Debye theory, 309-317
for dilute electrolytes, 313-314
and transient effects, 314-315
and dielectric constant, 137, 313
and ionic charge, 136, 313
and ionic strength, 137, 314
mechanistic interpretation, 137, 139
and primary salt effect, 321
tabulation of, 138
and temperature, 137
and viscosity, 136
Dipole-dipole interaction, 180
Dissociation field effect (see Wien effects)
Dymond-Alder potential, 42-43
Eadie transformation, 146
Effusion of gases
and isotopic enrichment, 15
and mean free path, 30
Knudsen flow, 30
and mean speed, 14
and molecular flux, 13-14
Elastic collision dynamics
and collision energy, 41
and impact parameter, 41
and intermolecular potential, 41
and microscopic reversibility, 79
simplified model, collinear geometries
equivalent single particle, 327
kinetic energy, 326
trajectories, 40-41
and velocity equilibration, 6
and walls, 7
Electric dipole transition moment, 167
Electrocapillarity, 222
Electrochemical potential, 55
Index
Electrolyte diffusion coefficient, 56
Electronic energy
and radiationless conversion, 173-174
tabulation of, 196
Electrophoretic field
in high electric fields, 66
for moderately dilute electrolytes, 63
Encounter complex, 136
Encounters
distinction from collisions, 136
rate of, 136, 309-311
Energy
of activation (see Arrhenius activation
energy)
of reaction, 100
Enthalpy
of activation, 303
estimation methods, 306
tabulations, 304
of reaction, 10 1
of solvation, 49
Entropy of activation, 303
estimation methods, 306
interpretation of
for gaseous reactions, 304-306
for solution reactions, 306
tabulation, 304
Enzyme catalysis (see Catalysis)
Equivalent conductance (see also Ion
equivalent conductance)
apparent
additivity rule, 51
concentration dependence, 51-52
concentration dependence, 52, 62-64
and electrophoretic field, 62-63
and ion atmosphere, 62-64
and relaxation field, 62, 64
Error function, 11, 21, 28
Eu(II), reaction with Fe (III), 94-96
pH dependence, 140-141
Excimers, 181
Excitation function, 264
Excitation transfer
in benzophenone, 197
diffusion limitations, 181
and excimer formation, 181
and molecular complexity, 180
and photodissociation, 181
and spin, 180
Explosions, 123
population, 205-209
Index
Explosions (cont.)
population (cont.)
explosion limits, 207
and surface effects, 205
thermal, 133, 202-204
diathermal coupling, 204
and heat evolution, 203
and surface effects, 204
Fast reaction methods, 83-85 (see also
Competition methods, Flash
photolysis, Rapid mixing methods,
Temperature jump)
Fe(ln
reaction of 4,7-(CH3h-1,IO-phenanthioline complex with IrCI/-, 106-107
reaction with Hg(II), 109
Fe(III)
reaction with bipyridine, 160
reaction with Cu(I), 160
reaction with Eu(II), 94-96
pH dependence of, 140-141
reaction with HCr04-, 109
reaction with V(III), 141-142
Feedback mechanisms
and branching chains, 206
and e1ectrocapillarity, 222
and glycolysis, 225
and thermal explosion, 202
and thermochemical oscillations, 223
Fick's laws, 25, 71
first law, 25
molecular justification, 33-34
second law, 27
First-order reactions
analysis of consecutive reactions, 114120, 153-156
and formation of trace intermediates,
1I5-1I6, 1I8
and steady-state hypothesis, II 6
and time-scale separation, 115, 117
Flames, 185
Flash photolysis
formation of trace intermediates, 83
nanosecond, 84
picosecond, 85
Flooding (see Isolation methods)
Fluctuations
and chemical hysteresis, 228
and nucleation, 228
and chemical oscillations, 211, 220
and thermal explosions, 204
361
Fluorescence, 170-172
quenching in diffusion-limited reaction,
315
vibrational, 185
Fourier's law, 24, 25, 71
molecular justification, 33
Franck-Condon principle
and dissociation, 174
and isomerization, 177
and shape of absorption band, 168
Friction, molecular, 57
Fumarate, reaction to I-malate, 161
r-function, 8
Gibbs free energy of activation, 303
Glutamate, reaction with pyridoxal phosphate, 104-105
Glycolysis, 224-225
IH
A-factor for reaction with h, 283-285
product distribution in beam reaction
with 2H2, 348-349
2H, beam reaction with Ch (see Ch)
3H
beam reaction with 2H 2, 249-250
reactions with propane, 250-251
H2
reaction with BCh, 190-191
reaction with Br2, 72-73, 77, 131-133
reaction with Ch, 133
reaction with lz, 72-73, 103
reaction with 02, 205-209
IH2H, isotope effect in beam reaction with
Kr+, 345-348
2H2
beam reaction with Ar+ (see Ar)
beam reaction with C, 248
beam reaction with 3H, 249-250
product distribution in beam reaction
with IH, 348-349
H/, beam reaction with He, 343-344
H3
potential-energy surface, 336
reaction profile, 274
Halogen hydrolysis, 110
Hanes transformation, 146
Harpooning model, 249, 262
2H3BPF 3, laser-induced decomposition,
189-190
H2C03, dissociation equilibrium, 67
362
Index
127-128
Irreversible thermodynamics
and reaction rate laws, 76
reciprocal relations, 26
and transport, 26
Isolation methods, 87-89
Index
363
Isotope separation
via effusion, 15
laser-controlled, 192-195
effect of pressure, 194
effect of vibrational anhannonicity,
194
364
Molecular diameter (cont.)
and reaction cross section, 236
in RRKM theory, 299-300
from transport data, 38-39
Molecular flux
and determination of pressure, 7
and effusion, 13-14
in molecular beams, 238
speed dependence in nonequilibrium
systems, 46
and transport, 31-32
Molecularity of reaction, 77
Molecular orbitals
in collision reactions, 258
and orbital symmetry correlations,
186-187
Molecular symmetry, and selection rules,
168
Molecular trajectories (see Elastic, Inelastic,
Reactive collision dynamics)
Multiple-photon absorption
and Arrhenius activation energy, 190
in laser-induced isotope separation, 194
and reactivity, 189
Multiple steady states, 122
analogy to van der Waals isotherms,
228-229
and chemical hysteresis, 226-230
and marginal stability, 227
Na
beam reaction with Br2, 265
beam reaction with CH3I, 265
Naphthalene, as excitation transfer donor,
181
Nernst-Einstein relation, 53-56
NO, as catalyst in reaction of 02 and S02,
142
N02, photolysis, 182
N205
decomposition of, 86
reaction with 03, 144, 161
Noncrossing rule, 187
Nonstationary states, 173
0, beam reaction with Br2 (see Br2)
O2
Index
Ohm's law, 25, 50, 71
deviations from, 65-67
Onsager conductance formula, 64-65, 68
Onsager reciprocal relations, 26
Orbital symmetry correlation, 185-188
Oscillatory reaction, 122
concentration waves in, 220
conditions for, 212
damping in closed systems, 213, 215
and detailed balance, 213
and distance from equilibrium, 212-213
and electrocapillarity, 222
and glycolysis, 224-225
and limit cycles, 214
and periodic precipitation, 221
and thermal oscillations, 223
Ostwald dilution law, 67
Periodic precipitation, 221-222
Perturbation methods (see Flash photolysis,
Relaxation methods)
Perylene, excitation transfer from
l-chloroanthracene, 198
Phase trajectory analysis, 212-214
Phosphorescence, 170-171
Photodissociation, 174-176
and chemiluminescence, 184
and excitation transfer, 181
Photoisomerization, 177-178
and orbital symmetry correlation,185188
Photolysis, 181-183
Photoproducts, methods of characterization, 168
Physical adsorption
energetics, 150
of rare gases, 150
and specificity, 150
Piperazine, reaction with pyridine-4aldehyde, 92-93
Poiseuille's law, 25, 71
molecular justification, 32
Polarography, 84
Potential-energy surfaces, 270-275
dimensionality, 271
for H3, 336
and isotopic substitution, 288
in many-atom systems, 275
models
for inelastic collisions, 329-334
for reactive collisions, 335, 340, 342,
348-350
and reaction profile, 271
saddle-point structure, 277, 282
Index
365
366
Rate law (cont.)
elementary, and principle of mass action,
73-76
experimental methods for establishing,
80-93
and linear irreversible thermodynamics,
76
and reaction complexity, 77
Reaction channels, 240
and Arrhenius A-factor, 244
energy dependence, 250-251
and fragmentation, 250
Reaction coordinate (see also Reaction
profile)
identification of, 282
as separable degree of freedom, 276-277
in unimolecular reaction, 291, 294-295
Reaction cross section
for atom-molecule reaction, 248-252
and collision energy, 245, 247, 249
and energy barrier, 245, 248-249
hard-sphere model for, 240-242
and internal states, 239, 261
for ion-molecule reaction, 245-248
and rate constant, 234-236, 239
as a thermal average, 240
Reaction order, 77
relation to decay constants, 114-120
Reaction probability, 235, 270
Reaction profile, 271
in barrierless reactions, 274-275
for H3 system, 274
and thermal reaction, 273
Reactive collision dynamics
effects due to noncollinearity, 345-348
multiple collisions, 345
model for collinear geometry
endoergic reaction, 342-345
energy partitioning, 338
exoergic reaction, 339-342
mass effects, 337, 341
thermoneutral reaction, 335-338
vibrational energy effects, 337
realistic model calculations, 348-351
Relaxation field
at high electric fields, 66
in moderately dilute electrolytes, 64-65
Relaxation methods, 83-84
decay constant in, 90
for coupled reactions, 91, 123
experimental limitations of, 91-92
in in vivo experiments, 93
Index
Relaxation time (see Decay constant)
Repulsive excited states, 175-176, 181
Resonance radiation, 170-171
Response time, 81
Rh2(OAc). 2H20, reaction with 5'adenosine monophosphate, 96-98
Rotational-vibrational energy transfer, 295
RRKM theory, 293-298
critique of, 299-302
random access assumption, 294, 300-301
rate constants from, 297-298
rotational-vibrational energy transfer
in, 195
strong collision assumption, 294
Runge-Kutta integration methods, 122
Scattering, reactive
complex formation, 252, 259-261
lifetime for, 259, 261
and long-range forces, 261
direct, forward-peaked, 252, 254-256
energy dependence, 254
and spectator stripping model, 255-256
measurement methods, 237-239
mUltiple channels, 244
product analysis, 239
rebound, 252, 257-258
and collision energy, 257
and exoergicity, 257
size of cross section, 257
transformation to center-of-mass coordinates, 252-253
accuracy limits, 253
velocity analysis, 239
Scavengers, in isotopic separation, 191
Second sound, 29
Selection rules for radiation
and molecular structure, 168
and spin, 167
vibronic effects, 168
Sensitization (see Excitation transfer)
SF6, isotopic separation of, 192, 194
Shock tubes, 82-83
Smog, photochemical, 182
S02, reaction with 02, 143
s20l-, reaction with H202, 103-104
Solvent cage
and encounter complexes, 136
and unimolecular reaction, 135, 139
Specific conductivity, 50-51
Spectator stripping model, 255-256
and collision energy, 256
Index
Spectator stripping model (cont.)
and impact parameter, 256
and intermolecular forces, 256
Spin multiplicity
and excitation transfer, 180
and selection rules, 167
Statistical thermodynamics, 316-318
Steady-state hypothesis
analysis of, in coupled first-order reactions, 116-120
in catalysis, 144-145
in chain reaction, 132-134
and branching, 134
and fluctuations, 204
in coupled arbitrary-order reactions,
120-122
in diffusion-limited reactions, 3 I I
transient effects, 314-315
in photochemistry, 172
in systems far from equilibrium, 210
steady state as a singular point, 2 I I, 2 I 9
and time-scale separation, 116-117, 120
Steric factor
in diffusion-limited reactions, 312
in hard-sphere model of reaction, 266
Stern-Volmer plot, 178
Stoichiometry
and mechanism, 72-73
and rate law, 72
Stokes' law
and ion equivalent conductance, 57-60
and stick/ slip boundary conditions, 57
Stopped flow, 82
Sucrose, inversion reaction, 161
Surface chemistry, 148-152
LEED studies, 150, 152
surface specificity, 150
Temperature jump, 83
in carbonyl addition reaction, 92-93
in reaction of Rh2(OAc). 2H20 and
5'-AMP, 96-98
Thermal conductivity, 25
of dilute hard-sphere gases, 32-38
and internal energy transport, 36, 46
and kinetic energy transport, 34, 46
Thermal diffusion, 26
Thermal oscillations, 223-224
Threshold energy
and Arrhenius activation energy, 242, 281
and pass energy, 273
Time-reversal invariance, 79-80
367
Time-scale separation
in arbitrary multistep processes, 120
in consecutive first-order reactions, 115
Trace intermediates (see Intermediates,
trace)
Trajectory analysis
for hard-sphere potential, 40-4 I
for Lennard-Jones potential, 40-41
in RRKM theory, 300
Transition state (see Activated complex)
Transmission coefficient, 278
in classical mechanics, 279-280, 282
isotope effect on, 289
in quantum mechanics, 279-280
Transmission function, 280
Transport coefficients
diffusion coefficient, 24
for dilute hard-sphere gas
interdependence, 35
qualitative properties, 34-35
tabulation, 36-38
temperatnre dependence, 37-39
theory of, 32-35
and reaction cross section, 236
and RRKM theory, 299-300
shear viscosity, 25
specific conductivity, 50-51
thermal conductivity, 24
Transport equations
force-flux relations
current flow, 50
diffusive flow, 24-25
heat flow, 24
viscous flow, 25
molecular justification, 29-34
partial differential equations
formulation, 26-27
solution of boundary value problems,
28,45
Tritium reactions (see 3H)
Tunneling, 273
Turnover number, 145
UF6, isotopic separation, 192, 195
Ultrasonic methods, 83
Unimolecular reaction-gases
and chemical activation, 30 I
energy dependence of rate constant,
291-293
phenomenological theory, 124-126
relation to radical recombination, 130-I 31
RRKM theory (see RRKM theory)
Index
368
U nimolecular reaction-gases (cont.)
simplified dynamical model for, 330-331
Slater theory, 301
Unimolecular reaction-solution
and diffusion, 135
and solvent activation, 135
and solvent cage, 135
V(II), reaction with [Co(NH3)5]20z'+' 160
V(III)
reaction with Fe(III), 141-142
reaction with Hg(II), 160
van der Waals forces, in adsorption, 150
Velocity distribution
in nonequilibrium systems, 29
positional dependence, 30, 40
specification precision limits, 2, 14
Vibrational energy, and reactivity, 190,261,
343-344
Vibrational relaxation, 169-171
Vibrational stabilization, 127
Vibrational-translational energy transfer
and chemiluminescence, 184
and laser activation, 189
and molecular complexity, 170
and pressure, 191
pro bability per collision, 170
and reactivity, 337, 340-341