ArticlePDF Available

Abstract and Figures

The concept of a rate-determining step (RD-Step) is central to the kinetics community, and it is basic knowledge even for the undergraduate chemical student. In spite of this, too many different definitions of the RD-Step appear in the literature, all of them with drawbacks. This dilemma has been thoroughly studied by several authors in the attempt to "patch" the drawbacks and bring the RD-Step to a correct physical meaning. Herein we review with simple models the most notable definitions and some challengers of the RD-Step concept, to conclude with the deduction that there are no rate-determining steps, only rate-determining states.
Content may be subject to copyright.
DOI: 10.1002/cphc.201100137
The Rate-Determining Step is Dead. Long Live the Rate-
Determining State!
Sebastian Kozuch*[a] and Jan M. L. Martin[b]
1. The Rate-Determining Step Dilemma
Contrary to other areas of human intellectual endeavor, in sci-
ence there is no place for ambiguity. A definition of a concept
must be (ideally) clear, unique and unequivocal; building those
definitions is the raison dÞtre for the existence of the IUPAC
body.[1] When a concept definition loses precision, or grows to
an unmanageable series of ad hoc conditions, we may be look-
ing at a case of an imperfect model of the scientific reality.
Some examples of this are the concepts of Lewis structures,
the formal charge, or the full set of “ideal” systems (ideal
gases, ideal solution, ideal covalent bond, etc). Sometimes, as
in the previous examples, this is only a philosophical matter, as
they are useful models. If we need more accuracy, we can add
a correction term or apply a more complex model. However,
sometimes the ambiguity in a definition can be a symptom of
a very pragmatic problem, which may lead to faulty conclu-
sions.
In chemical kinetics a central concept is the rate-determining
step (RD-Step), also called rate-controlling or rate-limiting
step.[1, 2] Herein we show with simple examples that the defini-
tions of this concept are not clear, unique or unequivocal. The
RD-Step description mutated from a clear but imprecise (and
frankly wrong) wording, to an obscure mathematical terminol-
ogy. These are symptoms of a pragmatic problem:[3–11] there
are no rate-determining steps![3, 4, 12] This statement has pro-
found consequences in homogeneous[3, 4, 13, 14] , heterogeneous
and enzymatic[6, 7, 15, 16] catalysis, kinetic isotope effect stud-
ies,[2,6, 8, 17] and the whole body of kinetic and mechanistic
chemistry.
2. Rate-Determining Step Faulty Definitions
Several definitions for the RD-Step can be found in the litera-
ture. We will cite the ones we consider most famous or promi-
nent, while the reader is invited to review the rest.[5, 18–23]
2.1. The Slowest Step of the Reaction[24–28]
The most typical definition of the RD-Step is “the slowest step
of the reaction”, always regarded as the “bottleneck”. In most
texts this definition goes side by side with the description of
the steady-state approximation (i.e. all steps in a multistep re-
action occur at the same rate when having unstable intermedi-
ates). This is a paradox. If we consider that the steady-state
regime is a very good approximation, then there is no slower
or faster step, and consequently there is no RD-Step ![2, 5, 8, 9, 18, 19]
2.2. The Step with the Smallest Rate Constant[29–31]
Sometimes the RD-Step is more strictly defined as “the step
with the smallest rate constant (k)” (strangely, this definition
was occasionally considered a synonym with the previous
one[29,30]). This definition is equivalent to saying “the step with
the highest activation energy”. It is a measurable designation,
thus more unambiguous and physical. However, unambiguous
and physical does not mean correct. A simple example can
show why.
Let us have a two-step reaction in the steady-state regime,
such as the one in Equation (1), with k1the smallest rate con-
stant:
C1G
k1
!k1
HC2
k2
!!C3ð1Þ
The concept of a rate-determining step (RD-Step) is central to
the kinetics community, and it is basic knowledge even for the
undergraduate chemical student. In spite of this, too many dif-
ferent definitions of the RD-Step appear in the literature, all of
them with drawbacks. This dilemma has been thoroughly stud-
ied by several authors in the attempt to “patch” the drawbacks
and bring the RD-Step to a correct physical meaning. Herein
we review with simple models the most notable definitions
and some challengers of the RD-Step concept, to conclude
with the deduction that there are no rate-determining steps,
only rate-determining states.
[a] Dr. S. Kozuch
Department of Organic Chemistry
The Weizmann Institute of Science
IL-76100 Rehovot (Israel)
Fax: (+972) 8-934-414 2
E-mail: sebastian.kozuch@weizmann.ac.il
[b] Prof. J. M. L. Martin
Center for Advanced Scientific Computing and Modeling (CASCAM)
Department of Chemistry, University of North Texas
Denton, TX 76203-5017 (USA)
ChemPhysChem 2011, 12, 1413 1418 "2011 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim 1413
The rate of reaction for this system is [Eq. (2)]:
r¼k2%k1
k!1þk2
C1
½( ð2Þ
If we consider k!1smaller than k2, then the rate of the reac-
tion will be [Eq. (3)]:
r¼k1C1
½( ð3Þ
Here the RD-Step definition being considered is correct. The
rate of the reaction depends solely on k1(the smallest rate
constant) and neither k2nor k!1influence the kinetics.
On the other hand, if k!1is bigger than k2, the rate equation
is [Eq. (4)]:
r¼k2%k1
k!1
C1
½( ð4Þ
In this case, how can we say that k1determines the rate of
reaction, when all the other rate constants have the same
weight? Furthermore, taking into account that k1/k!1=Keq1,
Equation (2) can be written as Equation (5):
r¼k2Keq1 C1
½( ð5Þ
and k2may be regarded as rate-determining, even though k1is
the smallest rate constant. Clearly the “smallest rate constant”
definition is flawed (especially when having reversible
steps[9, 32, 33]).
2.3. The Step with the Highest-Energy Transition State[34–37]
For the examples of the previous section [Eqs. (3) and (5)] the
“highest-energy transition state” definition will indeed work.
However, this is hardly a universal scenario. In a non-steady
state, two-step reaction of the type given by Equation (6):
C1
k1
!!C2
k2
!!C3ð6Þ
with an energy profile like that in Figure 1, the highest transi-
tion state is T1. However, the first step has smaller activation
energy compared to the second step (T1!I1<T2!I2, or equiva-
lently k1@k2). C1will be swiftly converted to C2, and then it will
slowly “drip” to C3. In this case the rate of reaction is [Eq. (7)]:
r¼k2C2
½( ð7Þ
The previous definition of RD-Step (“The step with the small-
est rate constant”) can work here, but the “highest transition
state” definition is erroneous (especially for irreversible
steps[2, 32]). In Section 4 (catalytic reactions), we show that even
for steady-state reactions this definition can be misleading.
2.4. The Step with the Rate Constant that Exerts the Stron-
ger Effect[1, 2, 6, 8]
IUPAC’s definition says that “a rate-determining step is an
elementary reaction the rate constant for which exerts a
strong effect—stronger than that of any other rate constant—
on the overall rate”.[1] In other words, a RD-Step is the one
whose rate constant is determining. This definition is a tautolo-
gy (a reiteration of the same idea in different words), but is ac-
tually the most accurate definition. In IUPAC’s gold book[1] a
mathematical description is added: the RD-Step is the one
whose control function (CF,[8] also called degree of rate con-
trol[20,21, 38] , and based on the similar concept of the sensitivity
function[6]) is the highest [Eq. (8)]:
CFi¼ki
z}|{
Normalization
factor
r%@r
@ki
|{z}
Effect of the
rate constant
on the rate
of reaction
¼@ln r
@lnkið8Þ
This differentiation must be computed maintaining all other
rate constants (kj,j¼6i,!i) and the equilibrium constant of that
step (Keq,i
) fixed. It is easier to understand these restrictions if
we pass from the k-representation (based on rate constants) to
the E-representation (based on energies)[3, 4, 7] according to
Equations (9) and (10):
Keq;i¼e!DGr;i=RT ð9Þ
ki¼kBT
he!DG!
i=RT ðTransition State TheoryÞð10Þ
where DGr,i is the Gibbs reaction energy of step i, and DGi
!is
the Gibbs activation energy of the step, as seen in Figure 2.
If we want to calculate the control function [Eq. (8)] for k1in
the two-step reaction of Equation (1), we have to calculate the
response of the global rate of reaction (r) to an infinitesimal
change in k1, keeping k2and Keq1 constant (Figure 2). According
to Equations (9) and (10), this means we must test rwith an in-
finitesimal change on T1(the Gibbs energy of the first transi-
tion state). In other words, the control function that is sup-
posed to test a rate-determining step, is actually examining a
rate-determining transition state.
Let us consider the case of Equation (5), where k2is the “RD-
Step” according to the control factor. In this case, a change in
Figure 1. Energy profile of a model non-steady state reaction with two irre-
versible steps.
1414 www.chemphyschem.org "2011 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim ChemPhysChem 2011, 12, 1413 1418
S. Kozuch and J. M. L. Martin
Keq1 has the same weight as a change in k2. A change in Keq1
keeping constant k2and k!1implies a change in k1, as appears
in Figure 3. So, we can make a change in k1that will affect the
rate of reaction as much as a change in k2, in spite of having a
null control factor for k1. How is this possible? [Eq. (11)]:[39]
r¼k2
z}|{
Step 2
is rate!
determining
Keq C1
½(¼k2%k1
z}|{
But step 1
is equally
important
k!1
C1
½(
ð11Þ
The problem resides in the definition of the control factor
(hence a problem in the definition of the RD-Step). The con-
straint of keeping Keq1 fixed makes this analysis impossible, but
lifting this constraint makes it impossible to define an unequiv-
ocal control factor. The reason is that the control factor can
only analyze the response to a transition-state change, and in
this case we are changing an intermediate (this issue was tack-
led in different ways by the “degree of turn-over frequency
(TOF) control”,[3,40] the “thermodynamic rate control for inter-
mediates”,[41, 42] the “thermodynamic control factor”,[43] the “sen-
sitivity”,[18, 44] and qualitatively by the Curtin–Hammett[45] and
Sabatier classical principles[46]).
Let us recapitulate the deficiencies of IUPAC definition of a
RD-Step:
1) The control factor, expressed as a function of rate con-
stants, has too many constraints (what we called ad hoc condi-
tions in Section 1). As graphically expressed in Figure 2, the
control factor is more a response to a transition state than to a
rate constant.
2) The control factor (and consequently this definition of the
RD-Step) cannot deal with valid cases that fall outside the con-
straint of the formula, as in Equation (11) and Figure 3. This
needs a new function that measures the response of the rate
with a change of intermediate energies.
3. Rate-Determining States: A Physically
Correct Concept
In Section 1 we argued that “when a concept definition loses
precision, or grows to an unmanageable series of ad hoc con-
ditions, we may be looking at a case of a bad model of the sci-
entific reality”. From Section 2 we can conclude that the RD-
Step is indeed a bad model of the scientific reality, as expressly
highlighted or hinted at by several authors. In that case, what
will be a good model for kinetic research ? As we saw in Sec-
tion 2.4, one transition state and one intermediate have the
potential to be the factors that shape the kinetics of the cycle.
In view of this, the correct concept should be the rate-deter-
mining states (RD-States), and not the rate-determining steps.
Can we build an unambiguous, clear, unique and unequivocal
definition for the RD-State? By all means.
Definition: Rate-determining states are the transition state
and intermediate which exert the strongest effect on the over-
all rate with a differential change on their Gibbs energies.
Mathematically, the transition state and intermediate with
the highest degree of rate control (XTS and XI) are the RD-States
[Eq. (12)]:[3,4, 40–42]
XTS;j¼1
r%@r
@!GTS;j
RT
"#
XI;j¼1
r%@r
@GI;j
RT
"#
ð12Þ
where GTS and GIare Gibbs energies of the intermediates and
transition states, and 1/ris a normalization factor. The degree
of rate control can be extended to study the sensitivity to con-
centrations.[12]
What is the reason that makes the RD-State a physically cor-
rect concept? The energy states are points in the reaction pro-
file space, while the steps are the processes that bind two con-
secutive points. In reality, the rate-determining zone[7] may
cover much more than two consecutive states, thus including
several steps (specifically all the steps that are between the
RD-States). This can be seen in Figure 3, where I1and T2are
the RD-States, and all the rate constants that are in between
(k1,k!1and k2) shape the reaction [Eq. (11)].
Figure 2. Energy profile of the two steps reaction of Equation (1). I1,I2and I3
are the Gibbs energies of the C1,C2and C3intermediates ; T1and T2the
Gibbs energies of the transition states. With Equations (9) and (10) we can
convert these energy states into rate and equilibrium constants. The control
function of k1[Eq. (8)] is measured by calculating the change in the rate of
reaction with a change in k1, maintaining k2and Keq1 constant. This is equiva-
lent to a change in T1.
Figure 3. A change in k1with fixed k!1and k2is equivalent to a change in I1,
the Gibbs energy of the first intermediate.
ChemPhysChem 2011, 12, 1413 1418 "2011 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.chemphyschem.org 1415
Long Live the Rate-Determining State !
4. RD-States in Catalysis
4.1. Introduction to Catalytic Cycles
Catalytic reactions are special in that they are cyclic: from the
catalyst’s perspective the initial and final state can be consid-
ered the same state, with the reactants and products as “addi-
tions”. A steady-state regime is easily achieved for most cata-
lytic cycles. In Figure 4 a model three-step catalytic cycle is de-
picted, both in the rate-constant (k) representation and the
energy (E) representation. Both representations are equivalent,
as we can convert each rate constant to an intermediate/tran-
sition state energy pair and vice versa, according to the transi-
tion state theory given by Equation (10).
In catalysis, the rate of reaction is called the TOF, the
number of cycles that occur in a specific time period. The reso-
lution of a three-step catalytic cycle in a steady-state regime is
beyond the scope of this text,[46, 47] and we write here only the
final formula (for simplicity we do not consider the concentra-
tion of reactants and products[4, 12, 47]) [Eq. (13)]:
TOF ¼
k1k2k3!k!1k!2k!3
k2k3þk3k1þk1k2þk!1k3þk1k!2þk!3k2þk!1k!2þk!2k!3þk!3k!1
ð13Þ
We can simplify the TOF equation by neglecting the lesser
terms. First, the product of forward rate constants (first term in
the numerator) is always bigger than the backward product
(second term) for exergonic reactions. Second, typically only
one of the nine terms in the denominator is significant; for a
reaction profile of the type of Figure 4, it will be k!2k1. In this
case, the resulting equation is given by Equation (14):
TOF ¼k1k2k3
k1k!2
¼k3Keq2 ð14Þ
Equation (14) shows that the TOF (the rate of reaction) is
completely independent of k1. However, not only is k1the
smallest forward rate constant, but T1is also the highest
energy state! Clearly, the rules of “smallest rate constant” and
“highest transition state” are not valid to establish the deter-
mining kinetic factors. Is there a simple way to find these de-
termining factors? There is, if we work in the E-representation
and try to find the RD-States instead of the RD-Step.
4.2. Visual Guide to Find the RD-States in Catalysis[3, 4, 7]
The net chemical flow in an exothermic reaction in the steady
state is always in the forward direction ; this fact will be the
first clue to find the RD-States in a catalytic cycle. The second
clue is that from each intermediate we can consider the effec-
tive energy barrier to be crossed as the highest energy transi-
tion state that occurs in the forward direction (this can be
thought as the Curtin–Hammett principle applied to catalytic
cycles[45]). The RD-States are the ones that have the highest ef-
fective energy barrier.
There is one factor that must be taken into account
when dealing with these systems. There is no starting or
ending point in a catalytic cycle, as the catalyst can theo-
retically continue to an infinite number of turn-overs. There-
fore, when we look for the highest TS following each inter-
mediate, we have to look further than one cycle. In Figure 5
we show the methodology to find the RD-States point by
point. Similar methods can be found elsewhere.[6, 15, 16, 32, 33, 43, 48]
Figure 4. A three-step model catalytic cycle, in A) the linear and closed k-
representations (rate-constant representations) and B) the E-representation
(energy representation). The kand Erepresentations are equivalent, and we
can translate one to the other with the transition state theory [Eq. (10)]. DGr
is the reaction Gibbs energy, independent from the catalyst.
Figure 5. Recipe to find the RD-States in a catalytic cycle: 1) Draw the Gibbs
energy profile of two consecutive cycles. 2) From each intermediate, find the
highest energy transition state in the forward chemical flow. 3) The deter-
mining states are the ones with the biggest span between the intermediate
in question and the following highest TS energy. These states are called the
TDI (TOF-determining intermediate) and the TDTS (TOF-determining transi-
tion state). The energy difference between the TDI and the TDTS is called
the energetic span (dE). In this example, I2is the TDI, and T1’ the TDTS. Note
that T1=T1+DGr.
1416 www.chemphyschem.org "2011 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim ChemPhysChem 2011, 12, 1413 1418
S. Kozuch and J. M. L. Martin
4.3. Mathematical Guide to Find the RD-States in Catalysis :
the Energetic Span Model
The catalytic profile of Figure 5 corresponds to the following
TOF equation [Eq. (15)] [a simplification of Eq. (13)]:
TOF ¼k1k2k3
k!2k!3
¼k1Keq2Keq3 ð15Þ
Almost all the rate constants appear in this formulation, thus
speaking of an RD-Step is a poor oversimplification (again).
From the visual technique of Figure 5, we found that I2and T1
are the RD-States. These two determining species are called
the TDI (TOF-determining intermediate, sometimes called the
resting state or the MARI[5, 11,23, 49]) and the TDTS (TOF-determin-
ing transition state, also called RD-Step !).[3] Can we find the
TDI and the TDTS from Equation (15) ? Yes, if we pass from the
k-representation to the E-representation by applying the transi-
tion state theory [Eq. (10)] to all the rate constants [Eq. (16)]:
TOF¼k1k2k3
k!2k!3
¼kBT
h
e!T1þI1%e!T2þI2%e!T3þI3
e!T2þI3%e!T3þI4
kBT
h
e!T1þI1þI2
eI4¼kBT
he!T1þI2!DGr
ð16Þ
In the last equality we considered that the starting point of
the second cycle is the end of the first one (I4=I1=I1+DGr).
As with the visual method of Figure 5, writing the rate equa-
tion in the E-representation naturally gave us I2and T1as the
RD-States. In this case DGris added as the acting TDTS is in
the second cycle (in Figure 5 it is shown as T1=T1+DGr). As a
rule, DGrappears in the TOF equation if the TDTS comes
before the TDI (when considering only one turn-over). After
this example, we can generalize the TOF expression in what is
called the Energetic Span Model [Eq. (17)]:[3, 4, 40]
TOF ¼kBT
he!dE
dE¼TTDTS !ITDI if TDTS appears after TDI
TTDTS !ITDI þDGrif TDTS appears before TDI
(ð17Þ
dEis the energetic span[50] of the cycle, and serves as the ap-
parent activation energy of the full catalytic cycle.[3, 4, 40, 43, 44, 48] It
depends directly on the RD-States (the TDI and TDTS), and has
no need of a cumbersome product of rate constants. Equa-
tion (17) is an excellent approximation when the degree of
TOF control of the TDI and TDTS is close to one (for the com-
plete TOF formula see references [3, 4, 12, 40]).
5. Conclusions
Herein we saw that the rate-determining step is a hard con-
cept to define, with each one of the definitions having defi-
ciencies. These are symptoms of a faulty concept : either we
must choose an RD-Step definition that fits the chemical prob-
lem,[2,39] or we are working in an unsuitable framework
(Figure 6). The literature is full of discussions and “patches” for
these deficiencies, pointing to a new viewpoint : the whole
idea of a RD-Step should be relegated to a historical place in
chemistry, and should make place for a concept with a stron-
ger physical basis, the rate-determining states.
Keywords: catalysis ·kinetics ·rate-determining step ·
reaction mechanisms ·transition-state theory
[1] The “IUPAC Gold Book” can be found under http ://goldbook.iupac.org.
[2] K. J. Laidler, Pure Appl. Chem. 1996,68, 149 192.
[3] S. Kozuch, S. Shaik, Acc. Chem. Res. 2011,44, 101 110.
[4] S. Kozuch, S. Shaik, J. Phys. Chem. A 2008,112, 6032 6041.
[5] G. Dj#ga-Mariadassou, M. Boudart, J. Catal. 2003,216, 89 97.
[6] W. J. Ray, Biochemistry 1983,22, 4625 4637.
[7] S. Yagisawa, Biochem. J. 1995,308, 305 311.
[8] K. J. Laidler, J. Chem. Educ. 1988,65, 250.
[9] R. K. Boyd, J. Chem. Educ. 1978,55, 84.
[10] P. W. N. M. van Leeuwen, Homogeneous Catalysis : Understanding the Art,
Kluwer, Dordrecht, 2004.
[11] P. Stoltze, Prog. Surf. Sci. 2000,65, 65 150.
[12] A. Uhe, S. Kozuch, S. Shaik, J. Comput. Chem. 2011,32, 978 985.
[13] S. Kozuch, S. E. Lee, S. Shaik, Organometallics 2009,28, 1303 1308.
[14] M. A. Carvajal, S. Kozuch, S. Shaik, Organometallics 2009,28, 3656
3665.
[15] P. E. M. Siegbahn, A. F. Shestakov, J. Comput. Chem. 2005,26, 888 898.
[16] P. E. M. Siegbahn, J. W. Tye, M. B. Hall, Chem. Rev. 2007,107, 4414 4435.
[17] D. B. Northrop, Biochemistry 1981,20, 4056 4061.
[18] J. A. Dumesic, G. W. Huber, M. Boudart in Handbook of Heterogeneous
Catalysis (Eds.: G. Ertl, H. Knçzinger, F. Sch$th, J. Weitkamp), Wiley-VCH,
Weinheim, 2008.
[19] M. Boudart, K. Tamaru, Catal. Lett. 1991,9, 15 22.
[20] C. Campbell, J. Catal. 2001,204, 520 524.
[21] A. Baranski, Solid State Ionics 1999,117, 123 128.
[22] J. Dumesic, J. Catal. 1999,185, 496 505.
[23] M. Vannice, Kinetics of Catalytic Reactions, Springer, New York, 2005.
[24] A Dictionary of Science, Oxford University Press, Oxford, 1999.
[25] S. S. Zumdahl, S. A. Zumdahl, Chemistry, Brooks Cole, Belmont, 2007.
[26] R. W. Missen, C. A. Mims, B. A. Saville, Introduction to Chemical Reaction
Engineering and Kinetics, Wiley, New York, 1999.
[27] C. Hill, An Introduction to Chemical Engineering Kinetics & Reactor Design,
Wiley, New York, 1977.
[28] P. Monk, Physical Chemistry: Understanding Our Chemical World, Wiley,
New York, 2004.
[29] R. Cammack, T. Atwood, P. Campbell, H. Parish, T. Smith, J. Stirling, F.
Vella, Oxford Dictionary of Biochemistry and Molecular Biology, Oxford
University Press, Oxford, 2006.
[30] J. Stenesh, Biochemistry, Plenum, New York, 1998.
[31] J. M. Berty, Experiments in Catalytic Reaction Engineering, Elsevier, Am-
sterdam, 1999.
[32] J. R. Murdoch, J. Chem. Educ. 1981,58, 32 36.
Figure 6. Which step is rate-determining ? Step 1, with the highest energy
transition state ? Step 2, with the smallest forward rate constant (highest ac-
tivation energy)? Or Step 3, having the determining transition state? There
are no rate-determining steps, but rate-determining states !
ChemPhysChem 2011, 12, 1413 1418 "2011 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.chemphyschem.org 1417
Long Live the Rate-Determining State !
[33] K. A. Connors, Chemical Kinetics : The Study of Reaction Rates in Solution,
VCH, Weinheim, 1990.
[34] E. V. Anslyn, D. A. Dougherty, Modern Physical Organic Chemistry, Univer-
sity Science, Sausalito, 2004.
[35] J. Hornback, Organic Chemistry, Thomson, Belmont, 2005.
[36] R. Hoffman, Organic Chemistry: An Intermediate Text, Wiley-Interscience,
New York, 2004.
[37] C. Suckling, Enzyme Chemistry: Impact and Applications, Chapman and
Hall, London, 1990.
[38] C. T. Campbell, Top. Catal. 1994,1, 353 366.
[39] M. Boudart, Top. Catal. 2001,14, 181 185.
[40] S. Kozuch, S. Shaik, J. Am. Chem. Soc. 2006,128, 3355 3365.
[41] C. Stegelmann, A. Andreasen, C. T. Campbell, J. Am. Chem. Soc. 2009,
131, 8077 8082.
[42] J. K. Nørskov, T. Bligaard, J. Kleis, Science 2009,324, 1655 1656.
[43] V. N. Parmon, React. Kinet. Catal. Lett. 2003,78, 139 150.
[44] H. Meskine, S. Matera, M. Scheffler, K. Reuter, H. Metiu, Surf. Sci. 2009,
603, 1724 1730.
[45] D. Y. Curtin, Rec. Chem. Progr. 1954,15, 111– 128.
[46] M. Boudart, Kinetics of Chemical Processes, Butterworth-Heinemann,
Boston, 1991.
[47] J. Christiansen, Adv. Catal. 1953,5, 311– 353.
[48] V. N. Parmon, React. Kinet. Catal. Lett. 2003,79, 303 317.
[49] M. Boudart, G. Dj#ga-Mariadassou, Kinetics of Heterogeneous Catalytic
Reactions, Princeton University Press, Princeton, 1984.
[50] C. Amatore, A. Jutand, J. Organomet. Chem. 1999,576, 254 278.
Received: February 18, 2011
Published online on April 27, 2011
1418 www.chemphyschem.org "2011 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim ChemPhysChem 2011, 12, 1413 1418
S. Kozuch and J. M. L. Martin
... The performance of RZABs were measured on a CHI 760E electrochemical workstation and 3 LAND-CT2001A test system. The specific capacity was calculated according to the equation ...
... (RDS), instead, there should be a "rate determining state". [3] That is, the catalytic activity should be co-determined by several steps. Based on such idea, one should avoid use ΔG max , but to build some newly proposed descriptor that abandon the using of rds. ...
Article
The decoration of γ-Fe 2 O 3 nanoparticles onto N,S co-doped carbon nanosheets enhances their catalytic activity for the oxygen reduction reaction (ORR) in both neutral and alkaline environments, making them a promising candidate for applications in various fuel cell scenarios.
... The values of the Tafel slope vary to a great extent among different measurements, leading to disparate reaction mechanisms reported in different studies. 20,21,[41][42][43][44] Furthermore, the Tafel slope exhibits a high sensitivity to adsorbate coverages. 1,21,45 Therefore, the RDS typically changes with electrode potential. ...
... Secondly, the coverage of adsorbates on the catalyst surface is assumed to be negligible. [20][21][22]42 In some cases, the second assumption is alleviated by determining the adsorbate coverages under quasi-equilibrium conditions. 21 However, the quasi-equilibrium conditions implied in both assumptions contradict with the fact that the reaction has a net rate, and that all elementary steps proceed with the same net rate. ...
Preprint
Full-text available
Electrocatalytic reactions, such as oxygen reduction/evolution reaction and CO2 reduction reaction that are pivotal for the energy transition, are multi-step processes occurring in a nanoscale electric double layer (EDL) at solid-liquid interfaces. Conventional analyses based on the Sabatier principle, using binding energies or effective electronic structure properties of the d-band center as descriptors, are able to grasp overall trends in catalytic activity in groups of catalysts. However, thermodynamic approaches fail to account for a plethora of electrolyte effects that arise in the EDL, including pH effects, cation effects, and anion effects. These effects have been observed to strongly influence electrocatalytic reactions. There is a growing consensus that the local reaction environment (LRE) prevailing in the EDL is the key to deciphering these complex and hitherto perplexing electrolyte effects. Equal attention is thus paid to designing appropriate electrolytes, positioning the LRE at center stage. Achieving this is essential for designing electrocatalysts with specifically tailored properties, which could enable much needed breakthroughs in electrochemical energy science. Theory and modeling are becoming increasingly important and powerful in addressing this multifaceted problem that involves physical phenomena at different scales interacting in a multidimensional parametric space. Theoretical models developed for this purpose should treat intrinsic multistep kinetics of electrocatalytic reactions, EDL effects from sub-nm scale to the scale of 10 nm, and mass transport phenomena bridging scales from < 0.1 to 100 μm. Given the diverse physical phenomena and scales involved, it is evident that the challenge at hand surpasses the capabilities of any single theoretical or computational approach. In this Account, we present a hierarchical theoretical framework to address the above challenge. It seamlessly integrates several modules: (i) a comprehensive microkinetic model accounting for various reaction pathways; (ii) an LRE model that describes the interfacial region extending from the nanometric EDL continuously to the solution bulk; (iii) first-principles calculations that provide parameters, e.g., adsorption energies, activation barriers and EDL parameters. The microkinetic model considers all elementary steps without designating an a priori rate-determining step. The kinetics of these elementary steps are expressed in terms of local concentrations, potential and electric field that are co-determined by EDL charging and mass transport in the LRE model. New insights on electrode kinetic phenomena, i.e., potential-dependent Tafel slopes, cation effects, and pH effects, obtained from this hierarchical framework are then reviewed. Finally, an outlook on further improvement of the model framework is presented, in view of recent developments in first-principles based simulation of electrocatalysis, observations of dynamic reconstruction of catalysts, and machine-learning assisted computational simulations.
... While a powerful concept for the definition of the catalytic activity has been derived by Kozuch in homogeneous catalysis (without applied bias), 34 their well-known free-energy span model has been transferred to electrochemical processes for energy conversion in the third generation of volcano plots. 2 The rate of a (chemical) catalytic cycle is governed by the free-energy span ranging from the intermediate with the smallest free energy to the transition state with the highest free energy in the freeenergy landscape (cf. Figure 3a). 34 Given that the CHE approach in conjunction with the concept of heuristic materials screening relies on the assessment of the intermediate states only, Exner suggested the free-energy span between the intermediate with the smallest and highest free energies, G max , as a measure of the electrocatalytic activity under equilibrium conditions (cf. Figure 3b). Subsequently, this analysis can be performed in a potential-dependent fashion by translating the free-energy landscape to overpotentials of practical interest (cf. Figure 3c), and the values of G max (η) for different materials can be analyzed to quantify activity trends. 2 The use of G max (η) reveals several advantages: first, this approach allows circumventing the combined methods of overpotential-dependent volcano plots, kinetic scaling relations, and ESSI by means of a more powerful descriptor that relies on fewer assumptions; additionally, there is also no requirement of an experimental parameter to consider the kinetics in the analysis. ...
Article
Full-text available
Conspectus Due to its importance for electrolyzers or metal–air batteries for energy conversion or storage, there is huge interest in the development of high-performance materials for the oxygen evolution reaction (OER). Theoretical investigations have aided the search for active material motifs through the construction of volcano plots for the kinetically sluggish OER, which involves the transfer of four proton–electron pairs to form a single oxygen molecule. The theory-driven volcano approach has gained unprecedented popularity in the catalysis and energy communities, largely due to its simplicity, as adsorption free energies can be used to approximate the electrocatalytic activity by heuristic descriptors. In the last two decades, the binding-energy-based volcano method has witnessed a renaissance with special concepts being developed to incorporate missing factors into the analysis. To this end, this Account summarizes and discusses the different generations of volcano plots for the example of the OER. While first-generation methods relied on the assessment of the thermodynamic information for the OER reaction intermediates by means of scaling relations, the second and third generations developed strategies to include overpotential and kinetic effects into the analysis of activity trends. Finally, the fourth generation of volcano approaches allowed the incorporation of various mechanistic pathways into the volcano methodology, thus paving the path toward data- and mechanistic-driven volcano plots in electrocatalysis. Although the concept of volcano plots has been significantly expanded in recent years, further research activities are discussed by challenging one of the main paradigms of the volcano concept. To date, the evaluation of activity trends relies on the assumption of proton-coupled electron transfer steps (CPET), even though there is experimental evidence of sequential proton–electron transfer (SPET) steps. While the computational assessment of SPET for solid-state electrodes is ambitious, it is strongly suggested to comprehend their importance in energy conversion and storage processes, including the OER. This can be achieved by knowledge transfer from homogeneous to heterogeneous electrocatalysis and by focusing on the material class of single-atom catalysts in which the active center is well defined. The derived concept of how to analyze the importance of SPET for mechanistic pathways in the OER over solid-state electrodes could further shape our understanding of the proton–electron transfer steps at electrified solid/liquid interfaces, which is crucial for further progress toward sustainable energy and climate neutrality.
... The energy span, dE, is the total energy barrier or kinetic resistor (kinestor) defined by the TOF determining intermediate (TDI) and the TOF determining transition state (TDTS), which is the rate-determining zone (RDZone). [59][60][61] We used Kozuch's AUTOF program to determine the TOF of each catalytic cycle. ...
Article
Full-text available
Magnesium‐based catalysts are becoming popular for hydroelementation reactions specially using p‐block reagents. Based on the seminal report from Schäfer's group (ChemCatChem 2022, 14, e202201007), our study demonstrates that the reaction mechanisms exhibit a far greater degree of complexity than originally presumed. Magnesium has a variety of coordination modes (and access to different hybridizations) which allows this electron‐deficient centre to modulate its catalytic power depending on the σ‐donor properties of the reagent. DFT calculations demonstrate several reaction channels closely operating in these versatile catalysts. In addition, variations in limiting energy barriers resulting from catalyst modifications were examined as a function of the Hammett constant, thereby predicting enhanced efficiency in reaction conversions.
... [19] Overall, in solution one reaction state in the conversion of 1 to 2 must be ratedetermining. [20] The fact that there is a monoexponential behavior in homogeneous solution makes it possible to determine the activation barrier of the reaction from 1 to 2 using an Arrhenius plot. Rate constants were determined at four temperatures (267 K, 283 K, 293 K and 300 K). ...
Article
Full-text available
The conversion of an azacalixpyridine‐supported Mo(0) tricarbonyl into a Mo(VI) trioxo complex with dioxygen (O2) is investigated in homogeneous solution and in a molecular film adsorbed on Au(111) using a variety of spectroscopic and analytical methods. These studies in particular show that the dome‐shaped carbonyl complex adsorbed on the metal surface has the ability to bind and activate gaseous oxygen, overcoming the so‐called surface trans‐effect. Furthermore, the rate of the conversion dramatically increases by irradiation with light. This observation is explained with the help of complementary DFT calculations and attributed to two different pathways, a thermal and a photochemical one. Based on the experimental and theoretical findings, a molecular mechanism for the conversion of the carbonyl to the oxo complex is derived.
Article
Full-text available
A series of linked-type 1,5-disubstituted tetrazoles (1,5-DS-Ts) were synthesised via IMCR and employed as synthetic platforms to access bound-type polyheterocycles. The sonochemical strategy include a double domino process under green conditions.
Article
Full-text available
Zinc–air batteries (ZABs) hold potential as clean, cost‐effective, and sustainable energy storage system for the next generation. However, the application of ZABs remains challenging because of their poor rechargeability and low efficiency . The design of efficient bifunctional catalysts toward oxygen reduction reaction (ORR) during discharging and the oxygen evolution reaction (OER) during charging is essential to developing rechargeable ZABs. Transition metal (TM)‐doped carbon (TM‐C) materials stand out from all the available bifunctional catalysts due to the excellent specific surface area, diverse morphological structures , and the multiple metal active sites formed after TM doping. This paper, therefore, focuses on the synthesis, electrochemical properties, and potential mechanism of TM‐C catalysts. To make a novelty and logical statement, the voltage difference (ΔE = Ei = 10 − E1/2) between the ORR/OER catalytic process is employed to categorize different TM‐C catalysts reported in recent years, which are divided into two groups: I (ΔE = 0.7 − 0.9 V) and II (ΔE = 0.5 − 0.7 V). The catalytic mechanisms of bifunctional catalysts are clarified. More ways and ideas for synthesizing high‐performance bifunctional TM‐C catalysts are also provided. Finally, the current problem and prospects of this group materials are presented.
Article
Stereoselective control of the cross metathesis of olefins is a crucial aspect of synthetic procedures. In this study, we utilized density functional theory methods to calculate thermodynamic and kinetic descriptors to explore the stereoselectivity of cross metathesis between allylbenzene and 2-butene-1,4-diyl diacetate. A ruthenium-based complex, characterized primarily by an anthracene-9-thiolate ligand, was designed in silico to completely restrict the E conformation of olefins through a bottom-bound mechanism. Our investigation of the kinetics of all feasible propagation routes demonstrated that Z-stereoisomers of metathesis products can be synthesized with an energy cost of only 13 kcal/mol. As a result, we encourage further research into the synthetic strategies outlined in this work.
Book
As the first edition of this book was going through the publication process, a revolution was taking place in the technologies available for the study of enzymes. The techniques of molecular biology, especially in genetic engineering of organisms and in site specific mutagenesis of genes, were established and were being brought into use to solve many problems in in enzymology. Added to these fundamental and applied science, not least advances the possibility of generating catalysts from antibodies has become a topic of major interest. These major innovations have changed the emphasis of much bioorganic research; whereas in the past, the protein was often the 'sleeping partner' ina study, its detailed function is now the major focus of scientific interest. Similarly in industry, the potential of genetically manipulated organisms to satisfy the needs for the production of chemicals and foodstuffs has been widely recognised. The second edition of 'Enzyme Chemistry, Impact and Applications' takes on board these new develop­ ments whilst maintaining the overall aims and views of the first edition. Many of the chapters have been completely rewritten to take account of advances in the last five years especially with regard to the impact of biologically based technologies. Although the book continues to approach its subject matter from the point of view of the chemist, the increased interdisciplinary content of much modern science will be obvious from the discussion.
Book
In the molecular sciences, enzyme chemistry occupies a special niche as one of the major contact points between chemical and biological disciplines. The special properties of enzymes as selective and efficient catalysts are so central to current challenges to chemists that the development of enzyme chemistry in the past thirty years has been a major stimulus to chemical research in general. On the one hand studies of the intrinsic properties of enzymes and, on the other hand, their applications to synthesis, drug design, and biosynthesis have had an immense impact. This book brings together in one volume essays describing several such fields with emphasis on the applications. It would be unnecessarily repetitious to outline the approach and contents of the book in a Preface; the first short chapter is more eloquent than a formal Preface can be. I shall therefore encourage you to begin with the Introduction in Chapter 1 and here I wish to extend my warm thanks to those who have contributed to the production of this book: the authors for their acceptance of the overall concept of the book and for the thoughtfulness of their writing; Dr Charles Suckling, FRS and Professor Hamish Wood for their constructive criticism of the whole book; and Dr John Buckingham and his colleagues at Chapman and Hall for their efficiency and enthusiasm in transforming the typescripts into the book that you now hold. Colin J. Suckling University of Strathclyde Contributors Donald H.
Article
This text is intended for an introductory course in bio­ metabolism concludes with photosynthesis. The last sec­ chemistry. While such a course draws students from vari­ tion of the book, Part IV, TRANSFER OF GENETIC INFOR­ ous curricula, all students are presumed to have had at MATION, also opens with an introductory chapter and then least general chemistry and one semester of organic chem­ explores the expression of genetic information. Replica­ istry. tion, transcription, and translation are covered in this or­ My main goal in writing this book was to provide stu­ der. To allow for varying student backgrounds and for pos­ sible needed refreshers, a number of topics are included as dents with a basic body of biochemical knowledge and a thorough exposition of fundamental biochemical con­ four appendixes. These cover acid-base calculations, principles of cepts, including full definitions of key terms. My aim has of organic chemistry, tools biochemistry, and been to present this material in a reasonably balanced oxidation-reduction reactions. form by neither deluging central topics with excessive de­ Each chapter includes a summary, a list of selected tail nor slighting secondary topics by extreme brevity. readings, and a comprehensive study section that consists Every author of an introductory text struggles with of three types of review questions and a large number of the problem of what to include in the coverage. My guide­ problems.
Book
The first comprehensive survey of the principles and applications of heterogeneous catalysis! With contributions from more than two hundred leading scientists, this book is indispensable for every scientist concerned with heterogeneous catalysis.
Article
This textbook contains all the information needed for graduate students or industrial researchers to design kinetic experiments involving heterogeneous catalysts, to characterize these catalysts, to acquire valid rate data, to verify the absence of mass (and heat) transfer limitations, to propose reaction models, to derive rate expressions based on these models and, finally, to assess the consistency of these rate equations. Langmuir, Freundlich, and Temkin isotherms are derived and the former are used in Langmuir-Hinshelwood (and Hougen-Watson) models, as well as reaction sequences without a rate-determining step, to obtain rate expressions on uniform surfaces. In addition, rate equations based on non-uniform (Temkin-type) surfaces are examined as an alternative approach. The most recent technique to calculate heats of adsorption and activation barriers on metal surfaces, the BOC-MP approach, is discussed in detail. Methods to measure metal surface areas and crystallite sizes using x-ray diffraction, transmission electron microscopy and various chemisorption techniques are discussed. Different experimental techniques to determine the influence of mass transfer limitations, especially within the pores of a catalyst, are reviewed in detail, with a particular emphasis on liquid-phase reactions. Many illustrations of these and other topics are provided along with numerous problems and a Solutions Manual for instructors. This book will be applicable to any graduate course in chemical engineering, chemistry or materials science that involves kinetics of catalytic reactions, including those catalyzed by enzymes. © 2005 Springer Science+Business Media, Inc., All rights reserved.