You are on page 1of 53

1

MiMiC REACTION KINETICS TUTORIAL

Extracting the Reaction Types and Parameters from Experimental Data

CONTENT
Primary Analysis of Kinetic Data
1. Introduction
1.1 Goals
2. Types of kinetic reactors
2.1 Open and Closed Reactors
3. Requirements to the Experiment.
4. Material Balances. Extracting the Net Rate of Change of Component
4.1. Material Balances
4.2. Steady-State and Dynamic Reactors
4.3. Transport in Reactors
4.4. Change due to the Chemical Reaction
4.5. Ideal Reactors
4.5.1 The Batch Reactor (BR)
4.5.2. The Continuous Stirred-Tank Reactor (CSTR)
4.5.3. The Plug-Flow Reactor
4.6. Ideal Reactors with Solid Catalyst
4.6.1. The Batch reactor
4.6.2. The Continuous Stirred-Tank Reactor
4.6.3. The Plug-Flow Reactor
4.6.4. The Pulse Reactor
4.7. Determination of the Net Rate of Component Change
5. Stoichiometry: Extracting the Reaction Rate from the Net Rate of Change of
Component.
5.1. Complexity and Mechanism. A Definition of Elementary Reactions
5.2. Homogeneous Reactions. Mass-Action-Law
5.3 Heterogeneous Reactions
5.4. Rate Expressions for Single Reaction. Stoichiometry
5.5. The Reaction Rates and Net Rate of Change of a Component: Relationships
5.6. Dimension of the Kinetic Parameters and Their Orders of Magnitude.
5.7. Distinguishing the Kinetic Dependences Based on Patterns
2

6. Influence of the Temperature on Kinetic Characteristics of the Single Reaction


6.1. Single Irreversible Reaction of the First Order
6.2. Single Reversible reaction.
6.3. Influence of the Temperature on the Equilibrium Composition of the First
Order Reaction.
6.4. Calculations of the Equlibrium Constant and reaction Quotient
6.5. Equilibrium and Optimum Conversion (Comparison)
7. Distinguishing the Reactions and Dependences
7.1 Distinguishing Single and Multiple Reactions
7.2. Distinguishing the Power Law and Non-Power Law Dependences.
7.3. Kinetic Decoding the Elementary Reactions from Complex Reaction
7.4. Four Types of Kinetic Scenarios

Notation
3

Primary Analysis of Kinetic Data.

1. Introduction.

Kinetic experiments are experiments in which the change of chemical composition of the
reacting mixture is monitored in time and/or as a function of different parameters of the system,
in particular inlet and initial chemical composition, temperature, catalyst state etc).

The main paradigm of contemporary chemical kinetics is the following: A chemical reaction is
complex and consists of elementary reactions for which the kinetic law is assumed to be known.

A kinetic analysis is focused on revealing this set and presenting corresponding mathematical
model (kinetic model)

1.1 Goals

Goals of the primary analysis of kinetic data are:

(a) Extracting the rates (rates of substance changes, rates of reactions etc) based on the data
of kinetic experiments, especially concentration measurements.
(b) Observing the coherency or incoherency between the observed kinetic dependences, and
categorizing the rate-parameters dependences
(c) Proposing the primary assumptions on the kinetic model and its mathematical description

Procedures for pursuing these goals depend on the type of kinetic experiments.

2. Types of Kinetic Reactors

2.1. Open and Closed Reactors

Kinetic experiments are performed in various types of reactors. Chemical reactors can be
classified as either open or closed reactors, depending on whether there is exchange of
matter with the surroundings. This classification has been adopted from thermodynamics,
in which a distinction is made between open and closed systems.

Closed reactors can exchange energy with the surroundings, but they cannot exchange
matter, while open reactors can exchange either matter and energy or only matter.
In chemical kinetics and engineering, the closed reactor is better known as batch reactor;
the open reactor as continuous-flow reactor.
4

Semi-open (or semi-closed) reactors also exist, in which only some type of material is
exchanged with the surroundings.
In pulse reactors, a small quantity of a chemical substance is injected into the reactor

3. Requirements to the Experiment

The primary information about chemical transformation in chemical reactors is a measurement of


chemical composition (concentration measurements)

The chemical processes occurring in laboratory reactors are complex and do not only consist of
chemical reactions but also physical phenomena, such as mass and heat transport. The major goal
of chemical kinetic studies is to extract intrinsic kinetic information related to the complex
chemical reaction. Therefore, the transport regime in the reactor has to be well-defined and its
mathematical description has to be reliable. We will use the latter as a measuring stick for
extracting the kinetic information. A typical strategy in kinetic experiments is the minimization
of the effects of mass and heat transfer on the rate of change of the chemical composition. In
accordance with this, the kinetic experiment ideally has to fulfill two main requirements:
isothermicity and uniformity of the chemical composition, which can be realized by e.g. perfect
mixing within the reaction zone.

A kinetic experiment should usually be performed under near isothermal conditions. The
temperature may be changed between two experiments. Temperature gradients across the reactor
can be minimized in various ways, for example by intensive heat exchange between the reactor
and the surroundings, by dilution of the reactive medium, or by its rapid recirculation.

Uniformity of the chemical composition at reactor scale is achieved by intensive mixing


using special mixing devices, internal, impellers, or external, recirculation pumps.

Both isothermicity and uniformity of the chemical composition can also be attained in
reactors in which the reaction zone is sufficiently small, e.g. differential plug-flow reactors
(PFRs), shallow beds etc.

4. Material Balances. Extracting the Net Rate of Change of Component

4.1. Material Balances

For any chemical component in any chemical reactor the material balance can be presented
qualitatively as follows:

temporal change of transport change due to



amount of component change reaction (1)
5

in which the temporal change of the amount of component, often termed accumulation, is its
change with respect to time at a fixed position, the transport change is the change caused by
motion of the component and the reaction change is the change caused by chemical reaction.

Equation (1) can be used for the classification and qualitative description of different types of
reactors for kinetic studies. The diagram presented below shows schematic representations of a
number of reactor types.

4.2. Steady-State and Dynamic Reactors

In dynamic reactors the temporal change of the concentration of a component, dci/dt 0, while in
steady-state reactors dci/dt = 0.

4.3. Transport in Reactors.

In laboratory kinetic reactors, the well-defined transport regime can be used as a measuring
stick for extracting the intrinsic kinetic dependences.

Examples of well defined transport regimes are pure convection or pure diffusion are.

For convection, the molar flow rate Fi (mol s-1) of a component i is determined as the product of
the total volumetric flow rate qV (m3s-1) and the concentration of the component ci (mol m-3):

Fi qV ci (2)
6

For diffusion, in the simplest case the molar flow rate of a component is determined in
accordance with Ficks first law:

dci
Fi Di A (3)
dz

where Di is the diffusion coefficient (m2s-1), A is the cross-sectional area of the reactor available
for fluid flow (m2) and z is the axial reactor coordinate (m).

In the model describing a batch reactor, the transport change term is absent based on the
assumption of perfect mixing.

Quite often the importance of transport phenomena has to be assessed at different scales,
that of the reactor being the largest. In heterogeneously catalyzed reactions the scale of the
catalyst pellets also has to be considered. The influence of inter- and intraparticle transport on the
reaction rate has to be eliminated experimentally and/or estimated quantitatively prior to the
kinetic experiments.

In perfectly mixed convectional reactors, the transport change can be represented as the
difference of convectional molar flow rates, qV 0ci 0 qV ci , where qV 0 and qV are the inlet and
outlet volumetric flow rates and ci 0 and ci are the inlet and outlet concentrations, or qV ci 0 ci if
qV qV 0 .

In purely diffusional reactors, the transport change in the simplest case can be
represented as the difference between diffusional flow rates in and out, Fi0 and Fi. Both flow
rates are written in accordance with Ficks first law:

ci ci
Fi 0 Di A ; Fi Di A (4)
z z z z z

Then,

c c 2ci
Fi 0 Fi Di A i Di A i Di A 2 z (5)
z z z z z z
7

4.4. Change due to the Chemical Reaction

A change due to reaction(s) is calculated as a function R of concentration (c) and temperature


(T), where R is the net rate of change of the component per unit reaction volume (mol m-3s-1).
In chemical kinetics and chemical engineering, the concept of fractional conversion, or simply
conversion, Xi, is widely used. The conversion of a component is defined as
Fi 0 Fi
Xi ; Fi Fi 0 1 X i
Fi 0 (6)
or, when the reaction volume is constant as
ci 0 ci
Xi ; ci ci 0 1 X i
ci 0 (7)

Xi is dimensionless and varies between 0 and 1.


For the simplest case, a first-order reaction, the reaction rate can be expressed as

r kci kci 0 1 X i
(8)

with k the reaction rate coefficient (s-1).

4.5. Ideal Reactors

In chemical kinetics, three reactors can be considered as typical or ideal reactors

First, lets analyze ideal reactors of constant reaction volume in which a stoichiometrically single
reaction takes place, without explicitly taking into account the presence of a solid catalyst, i.e.
we are assuming the reaction is not catalyzed or is homogeneously catalyzed. Reaction rates are
all expressed in moles per unit of reaction volume per second (mol m-3s-1). If solid catalysts are
involved it is more convenient to express reaction rates per unit mass or unit surface area of
catalyst.
Three mentioned ideal reactors:

(a) Batch Reactor (BR)


(b) Continuously Stirred Tank Reactor
8

(c) Plug Flow Reactor

4.5.1 The Batch Reactor (BR)

In an ideal batch reactor, i.e. a dynamic closed reactor with perfect mixing, the transport term is
absent, and Eqn. (1) becomes

temporal change of change due to



amount of component reaction (9)

In definition, the BR is the dynamic reactor.

The simplest mathematical model for the temporal change of any component in a batch reactor
of constant reaction volume is

dci
Ri i r (10)
dt

where Ri is the net rate of change of the component per unit reaction volume (mol m-3s-1), i is
the stoichiometric coefficient and r is the reaction rate (mol m-3s-1).

For a reversible reaction, the reaction rate is a combination of the rates of the forward and
backward reactions:

r= rf - rb (11)
We will analyze these topics in more detail in the Section Stoichiometry

The state in which dc/dt = 0, is called equilibrium. In this state rf - rb =0 and rf = rb

In the batch reactor, the net rate of component change is determined from the time
derivative of the reactant concentration, Ri dci/dt, see Eqn (10)

There are two cases:

(a) A reaction is performed in the gaseous phase: gas reactors. A volume of the reacting
mixture is considered to be constant (It is a volume of the reactor). A pressure of the
reacting mixture can be changed as a function of reactant conversion.
9

(b) A reaction is performed in the liquid phase: liquid reactors. A volume of the reacting
mixture can be changed as a function of reactant conversion. A pressure is typically
constant and equals to the atmospheric pressure.

4.5.2. The Continuous Stirred-Tank Reactor (CSTR)

A continuous stirred-tank reactor (CSTR) is an open reactor with perfect mixing (gradientless
reactor) and only convective flow. This mixing can be achieved not only by internal but also by
external recirculation. The material balance for any component in a non-steady-state CSTR can
be written as

dci q c q c
Ri V 0 i 0 V i (12)
dt V
with V, the reaction volume (m3).

Obviously, the batch reactor (BR) is a particular case of the continuous-stirred-tank-reactor


(CSTR), if there is no exchange by matter between the reactor and surroundings, e.g.

qV qV 0 , and equals 0.

Typically, the CSTR-experiments are performed in the steady-state regime (the steady-state-
reactor),

i.e. dc/dt = 0

At steady state, the net rate of change of component i can be determined from

qV 0ci 0 qV ci
Ri (13)
V

If qV qV 0 , Eq. (13) can be expressed as

qV 0 ci 0 ci c c
Ri i0 i (14)
V
10

where = V/qV0 is the space time (s). It is denoted as space time because its definition involves a
spatial variable, V, which distinguishes it from the astronomic time. It corresponds to the
average residence time in an isothermal CSTR.

Therefore, in the steady-state continuous stirred-tank reactor (CSTR), the net rate of
production of a component is equal to the difference between the corresponding outlet and
inlet concentrations, divided by the space time .

Ri = - (ci0 ci)/, see Eqn (14)

For the simplest case, i.e. the first-order reaction, considering see Equations (7) and (8), Equation
(14) can be written in terms of conversion as follows:

Xi
k 1 X i (6)

or

k
Xi (7)
1 k

The term k, also known as the first Damkhhler number DaI , i.e. the ratio of the time scale for
transport from inlet to outlet of the reactor to the time scale of the reaction, is the main
characteristic of the CSTR. For large k (>>1), conversion is complete, Xi = 1. If k is small
(<<1), Xi = k. Knowing the conversion, an apparent first-order rate coefficient can be
determined:

Xi 1
k (17)
1 Xi

4.5.3. The Plug-flow Reactor (PFR)


11

In an ideal plug-flow reactor (PFR), it is assumed that perfect uniformity is achieved in the
direction perpendicular to that of the flow, i.e. in the radial direction. This is relatively easy to
achieve in tubular reactors with high aspect ratio, i.e. with large length-to-diameter ratio. Axial
diffusion effects are also neglected. The composition of the fluid phase varies along the reactor,
so the material balance for any component must be made for a differential element:

dci
dV Ri dV qV dci (8)
dt

In a more rigorous form, Eq. (18) can be written as a partial differential equation:
c c
dV i Ri dV qV i dz (9)
t z

Using qV = uA and dV = Adz, where u is the superficial fluid velocity (m s-1), Eq. (19) can be
written as

ci c
Ri u i (10)
t z
or
ci ci
Ri (11)
t
with = z/u.

The typical model of PFR is related to the steady-state case, where ci/t = 0, and the model
equation for an ideal PFR can be expressed by the ordinary differential equation

dci
Ri (22)
d
12

which remarkably is identical to the expression for a batch reactor, Eq. (10). The only difference
is the meaning of the term time used. In the model for the batch reactor, the time is the time of
the experimental observation or astronomic time, whereas the time in the model for the plug-
flow reactor is the space time, .

To obtain the net rate of production in accordance with Eq. (22), we have to find the
dci
derivative Ri based on experimental data. Typically, for a steady-state PFR, the
d
longitudinal concentration profile is difficult to measure, and the measured characteristic is the
gas-phase concentration at the reactor exit. At the given reactor length, space times will be
changed be the systematic increasing or decreasing the value of the flow-rate. This is a so-called
differential method of analyzing PFR data. In practice, however, the experimental error is rather
high because net rates of changes of components cannot be measured directly, in contrast with
the CSTR data analysis.

In some cases, a model of the so called differential plug-flow reactor (differential PFR) is
applied. This type of reactor can be considered a hybrid between a CSTR and an integral PFR,
with conversions sufficiently small not to affect the reaction rate. This model can be applied for
reactors with thin active zone or a high feed flow rate or both. In a differential PFR, the reaction
zone can be assumed to be perfectly mixed. The concentrations inside the reaction zone can be
taken as the inlet or outlet concentrations or as the average of these concentrations. Eq. (22) for
the differential PFR is equivalent to Eq. (14) for the CSTR.

4.6. Ideal Reactors with the Solid Catalyst

4.6.1. The Batch Reactor

The equivalent of Eq. (10) for a batch reactor containing a solid catalyst is

1 b dci
RW ,i i rW (12)
cat 1 b dt
13

1 1
where RW,i is the net rate of change of component i per unit mass of catalyst ( mol kg cat s ), rW is
s ), b is the void fraction of the catalyst
1 1
the reaction rate per unit mass of catalyst ( mol kg cat
bed1 (-) and cat is the density of the catalyst pellet ( kg cat mcat
3
).

4.6.2. The Continuous-Stirred-Tank Reactor

In the case of a CSTR with a solid catalyst, the material balance, Eq. (12), can be written as

1 b dci q c q c
RW ,i V 0 i 0 V i (13)
cat 1 b dt Wcat

with Wcat the mass of catalyst in the reactor (kgcat).

The net rate of change of component i at steady state, i.e. dc/dt = 0, can be determined
from

qV 0 ci 0 ci F Fi
RW ,i i0 (14)
Wcat Wcat
With
Fi 0 Fi
Xi (15)
Fi 0
we obtain
Fi 0 X i
RW ,i (16)
Wcat

where W/Fi0 is also often referred to as the space time.

1
The void fraction is the ratio of the volume of fluid to that of the total reactor volume, Vf/V = Vf/(Vf + Vcat).Typical
values of the void fraction are placed in the domain [0.4; 0.46], see Ergun, S., CEP, 48 (2), 89-94 (1952)
14

4.6.3. The Plug-Flow Reactor

For the PFR with a solid catalyst, the equivalent of Eq. (18) is

1 b dc
dWcat i RW ,i dWcat qV dci (17)
cat 1 b dt
At steady state, i.e. dc/dt = 0
dci dFi
RW ,i qV (18)
dWcat dWcat
or

Fi 0 dX i dX i
RW ,i (19)
dWcat d Wcat Fi 0

Similar to the traditional plug flow reactor (see section 4.5.3), in some cases a model of the so
called differential plug-flow reactor (differential PFR) is applied. In the same way, the
concentrations inside the reaction zone can be taken as the inlet or outlet concentrations or as the
average of these concentrations. Eq. (30) for the differential PFR is equivalent to Eq. (27) for the
CSTR.

4.6.4 The Pulse Reactor

In a pulse reactor, which typically contains a fixed catalyst bed, a small amount of a component
is injected into the reactor during a small time interval. In a conventional pulse reactor, the
component is pulsed into an inert steady carrier-gas stream. The relaxation of the outlet
composition following the perturbation by this pulse provides information about the reaction
kinetics.

In the TAP reactor, created by John Gleaves in the 1980, no carrier gas stream is used
and the component is pulsed directly into the reactor. Transport only occurs by diffusion, in
particular Knudsen diffusion. In a thin-zone temporal-analysis-of-products reactor (TZTR), the
catalyst is located only within a narrow zone, similar to the differential plug-flow reactor.
15

In the industry, pulse regimes are applied in some adsorption/desorption processes, in particular
in some swing adsorption processes in which gas species are separated from a mixture of gases
under pressure according to the different affinity for an adsorbent material

The net rate of component change in the catalyst zone of the TZTR is the difference between
two diffusional flow rates at the boundaries of the thin active zone divided by the mass of
catalyst in the reactor:

Fi (t ) Fi 0 (t )
RW ,i (20)
Wcat

To some extent this is analogous to the case of the steady-state differential PFR and to the
steady-state CSTR, in which the reaction rate is given by the difference between convectional
flow rates.

4.7. Determination of the Net Rate of Component Change

Summing up, in typical cases the conceptual difference between the different methods of
measuring the net rate of change of component can be presented as follows:

In the dynamic batch reactor (BR), the net rate of component change is determined from the
time derivative of the reactant concentration, see Eqn (10)

Ri dci/dt, (32)

In the steady-state continuous stirred-tank reactor (CSTR), the net rate of production of a
component is equal to the difference between the corresponding outlet and inlet concentrations,
divided by the space time, see Eqn (14):

Ri = - (ci0 ci)/, (33)

Finally, in the plug-flow reactor (PFR), the net rate of change of component is
determined from the derivative of the component concentration with respect to the reactor length
length or reactor space time,

Ri dci/dz, which with = z/u can be written as


16

Ri dci/d (34)

In the PFR, the typical measured characteristic is the gas-phase concentration at the reactor exit.
At the given reactor length, space times will be changed be the systematic increasing or
decreasing the value of the flow-rate. This is a so-called differential method of analyzing PFR
data.

The conceptual material balance equation, Eq. (1), is often written as

temporal change of flow flow reaction



amount of component in out change (35)

In a batch reactor, both flow in and flow out terms are absent, while in a CSTR both flow
terms are present. In a PFR, both flow terms are present too, but flow in flow out is
presented in differential form. In pulse reactors, initially there is only the flow in term, while
later there is only the flow out term.

Experimentally, these three types of rate measurement differ regarding both the reactor
construction and the extraction of the Ri value. Technically, the PFR is the simplest reactor.
However, the CSTR has an advantage over the BR and PFR as for the accuracy of the rate
estimated: an experimental error in concentration is less than the error in concentration
derivative. Also the differential method of analyzing PFR data is a quite complicated
experimental procedure. Statistical problems in kinetic analysis will be discussed later.

5. Stoichiometry: Extracting the Reaction Rate from the Net Rate of Change of
Component.

5.1. Complexity and Mechanism. A Definition of Elementary Reactions

Most chemical reactions are complex in nature. They have so called a reaction mechanism or
equivalently mechanism or detailed mechanism that consist of many steps, referred to as
elementary steps.

The detailed mechanism is a combination of elementary steps. Each step consists of a forward
and a reverse elementary reaction. Rigorously, every step and every overall reaction is
17

reversible. However, in reality many steps and overall reactions can be considered to be
irreversible.

The IUPAC Gold Book definition of an elementary reaction is as follows:

An elementary reaction is assumed to occur in a single step and to pass through a single
transition state. (McNaught, A.D. and Wilkinson, A. (1997) IUPAC Compendium of Chemical
Terminology (Gold Book), 2nd, Blackwell Science, Cambridge, UK, 464 pp)

In other words, one elementary energetic barrier corresponds to one elementary reaction. An
elementary reaction takes place exactly as written.

The rate of an elementary reaction can be defined as the number of elementary acts of chemical
transformation per unit volume of the reaction mixture (or unit catalyst surface area, etc.) per
unit time.

If an elementary reaction involves one reactant molecule (A B), it is classified as a


unimolecular (monomolecular) reaction, or the first-order reaction.

If two molecules take part in the reaction

(e.g. 2A B or A + B C), the reaction is called a bimolecular reaction, or the second order
one. Finally, with the participation of three molecules (3A B, or 2A + B C), the reaction is
said to be termolecular, or the third order one. The simultaneous interaction of more than three
reactant molecules in one elementary reaction is believed to be highly improbable and even
termolecular reactions are very rare.

The reaction rate of an elementary step is determined by the difference between the rates
of the forward and the backward reactions (see Eqn (11))

r = rf -rb

where r, rf and rb are the rate of the step, the rate of the forward reaction and the rate of the
backward reaction, respectively. In many cases (see below), symbols + and - are used
18

alternatively for indicating the forward and backward reactions instead of symbolsf andb,
respectively

5.2. Homogeneous Reactions. Mass-Action-Law

Consider the general elementary step



k
1 A1 + 2 A2 1 B1 + 2 B2
(36)
k

or

i Ai
k
i Bi
(37)
k

where Ai and Bi are reactants and products with i and i the absolute values of their
stoichiometric coefficients and k+ and k- are the rate coefficients for the forward and reverse
reactions. In addition to the limitation on the values of i and i ( 3), the sum of coefficients i
must also not be larger than three.


means that both the overall reaction and every step, rigorously speaking,
The symbol

are reversible, i.e. consist of two reactions, a forward and a backward (reverse) one.

The dependence of the rates of the forward and reverse reactions on the concentrations of
reactants is expressed in terms of the mass-action law as

r k cA11 cA22 k cAii (38)

r k cB11 cB22 k cBii (39)

where cAi and cBi are the concentrations of reactants and products (mol m-3). The rate coefficients
determine the reaction rates of the forward and the backward reaction at unitary values of
19

reactant concentrations. They are governed by the Arrhenius dependences (40) and (41) and
increase exponentially with temperature:

E
k k0 exp a (40)
RT
g
E
k k0 exp a (41)
RT
g

Here k0 and k0 are the pre-exponential factors, Ea and Ea are the activation energies of the
forward and backward reactions, respectively; Rg is the universal gas constant, and T is the
absolute temperature. The ratio of the rate coefficients of the forward and reverse reaction
determines the equilibrium coefficient Keq:

k
K eq (42)
k

The difference between the activation energies for the forward and backward reactions
determines the reaction enthalpy:

r H Ea Ea (43)

For an exothermic reaction, in which heat is released, rH < 0.

For an endothermic reaction, in which heat is consumed, rH > 0.

An example of the monomolecular or first-order reactions is


20

C2H5Br
C2H4 + HBr (44)

and

CH3N2CH3
C2H6 + N2 (45)

The rates of these reactions are r kcC2H5Br and r kcCH3N2CH3 , respectively.

The reactions

2 NOI
2 NO + I2 (46)

and

CO + O2
CO2 + O (47)
are bimolecular or second order with rates r kcNOI
2
and r kcCOcO2 , respectively.

The reaction

2 NO + O2
2 NO2 (48)

is termolecular, or third order with r kcNO


2
cO2 .

As a rule, in elementary reactions reactants and products are different. An example of an


exception to this rule is one of the steps in the thermal dissociation of hydrogen:

H + H2


3 H (49)
21

In this step, one of the reactants (radical H) is also the product of the reaction. This is an
example of a so-called autocatalytic reaction.

5.3 Heterogeneous Reactions

In the case of heterogeneous gas-solid catalytic reactions, reactants in elementary steps can be
gas-phase components or surface intermediates. Similar to Eqn. (36), any chemical step that
involves a catalyst can be written as



k
i Ai + j Xj i Bi + j Yj
(50)
k

where Ai and Bi are reactants and products in the gas phase with i and i the absolute values of
their stoichiometric coefficients and Xj and Yj are surface intermediates with j and j the
absolute values of their stoichiometric coefficients.

Typically, Eq. (50) is of the form

A
k
+ j Xj B + j Yj
(51)
k

Moreover, and are either one or zero, i.e. it is assumed that in an elementary catalytic
reaction only one molecule from the gas phase reacts (e.g. CH4 + Z CH2Z + H2 ) or none at
all (e.g. CHOHZ COZ + H2). The stoichometric coefficients j and j are assumed to be
equal to one, two, or, rarely, three, and j and j are not larger than three.

The rates of the forward and backward reactions can be written similarly as Eqns. (38)
and (39):


rS k cAii Xjj (52)
22


rS k cBii Yjj (53)

in which i and i are either zero or one and j 2 if i = 1, j 3 if i = 0, j 2 if i = 1 and j


3 if i = 0. rS and rS are the forward and backward reaction rate per unit catalyst surface area

s ). X j and Y j are the normalized surface concentrations or fractional surface


2 1
( mol mcat

coverages of surface intermediates:

X j Yj
X j ; Y j (54)
t t

with X j and Y j the surface concentrations of species Xj and Yj and t the total concentration of
2
surface intermediates, including free active sites ( mol mcat ).

The total concentration of active sites can be determined in separate experiments, e.g. in
adsorption experiments or in multi-pulse response experiments under high vacuum conditions.

In the case of heterogeneous liquid-solid reactions, reactants in elementary steps can be liquid
components (components in the bulk of the liquid) or some intermediates on catalyst surface.
The reaction may occur within the film of certain depth on the liquid surface..

5.4. Rate Expressions for Single Reaction. Stoichiometry

For a system without exchange of matter with the surrounding medium, the rate of a
stoichiometrically single reaction, both elementary and complex, can be expressed as

1 dnAi 1 dnBi
r (55)
iV dt iV dt
23

Here nAi and nBi are the number of moles of reactants and products and V is the volume of the
reaction mixture. The rate r is expressed in mol m-3s-1.

More specifically, for a heterogeneous catalytic reaction in a closed system, e.g. a gas-
solid reaction, the reaction rate can be expressed as, for instance

1 dnAi 1 dnBi
rS (56)
i Scat dt i Scat dt
2
Here Scat is the catalyst surface area ( mcat ) and the rate rS is the rate per unit catalyst surface area
2 1
( mol mcat s ).

The reaction rate can also be expressed per unit volume of catalyst Vcat ( m3cat ), with rV in
3 1 1 1
mol mcat s , or per unit mass of catalyst Wcat (kgcat), with rW in mol kgcat s . These rates can be

easily transformed:

Vcat V W
rS rV cat cat rW cat rW (57)
Scat Scat Scat

3
with cat the density of the catalyst ( kg cat mcat ).

For chemical processes without a change in the number of moles during the course of the
reaction, the reaction rate takes the traditional form:

1 dcAi 1 dcBi
r (58)
i dt i dt
24

with concentrations in mol m-3.

Similarly, the rate of catalytic heterogeneous reaction can then be written as

Vf dcAi V dcBi
rS f (59)
i Scat dt i Scat dt

with Vf the volume of the fluid phase (m3).

5.5. The Reaction Rates and Net Rate of Change of a Component: Relationships

There is a big difference between the reaction rate and the net rate of change of a component.

For a single stoichiometric reaction step, the relationship between r and Ri can be
expressed as follows:

Ri
r or Ri i r (60)
i

where i is the stoichiometric coefficient of the chemical component i. The convention is to


assign negative coefficients to reactants and positive coefficients to products. Thus Ri is also
negative for reactants and positive for products.

For example, the reaction rate for the elementary step A =B is given by

r r r k cA k cB (61)
25

Now,

RA RB RA RB
r r RA RB (62)
A B 1 1

For the elementary step 2 A + B = 3 C,

r r r k cA2 cB k cC3 (63)

and r = RA/2 = RB = RC/3 or RA = 2r, RB = r and RC = 3r. (64)

Because of this difference, the definitions of the net rate of production of a component
and the reaction rate have to be carefully distinguished. The net rate of production of a
component is an experimentally observed characteristic. It is the change of the number of
moles of a component per unit volume of reactor (or catalyst surface, volume or mass) per unit
time.

The reaction rate r can be introduced only after a reaction equation has been assumed
with the corresponding stoichiometric coefficients. Then, the value of reaction rate can be
calculated based on the assumed stoichiometric equation using Eqs. (55) and (56).

This is a really big conceptual difference between the experimentally observed net rate of
production of a component, Ri, and the calculated reaction rate, r.

This difference between the two rates has to be taken into account even if we consider our
reaction as a single one, say an isomerization reaction of a reactant A into a product B. Even
knowing the rate of consumption of A, we will obtain different rates of reaction depending on
what kind of elementary reaction we are going to assume: A B or 2A 2B.

In the case that a component is participating in multiple reactions, Ri is a linear


combination of the rates in which this component is consumed or formed in the steps taking
26

place, rs. The coefficients in this linear combination are the stoichiometric coefficients is of the
component in each of the steps:

Ri ris is rs (65)
s s

The main methodological lesson is: Do not mix the experiment with its interpretation.
The net rate of production of a component is an experimentally measured value. The reaction
equation, on the other hand, is a result of our interpretation, and it can be written arbitrary.

Therefore, the reaction rate calculated in accordance with this reaction equation is a part of our
interpretation as well.

Both rates, the net rate of component change and the reaction rate, are used in many further data
procedures, i.e. determination of kinetic coefficients k, pre-exponential factors ko, energies of
activation Ea, kinetic orders etc

5.6. Dimension of the Kinetic Parameters and Their Orders of Magnitude.

The dimension of the rate coefficient k depends on the type of chemical reaction. In the case of a
homogeneous chemical reaction, i.e. a reaction involving a single phase, the dimension of k, [k],
is

k
r (67)
i
cAi
i
27

for the forward reaction of (38) and

k
r (68)
i
cBi
i

for the backward reaction of (39).

Table 1 shows the dimension of the rate coefficient for the three types of reactions. The
dimension of the pre-exponential factor is the same as the dimension of the rate coefficient k.

Table 1. Dimension of the rate coefficient k for homogeneous reactions


reaction type [k]

Monomolecular mol m 3s 1
3
s 1
mol m

Bimolecular mol m3s 1


m3mol1s 1
mol m 3 2

Termolecular mol m 3s 1
m6 mol2s 1
mol m 3 3

Table 2 shows orders of magnitude of the kinetic parameters for first-order reactions. The pre-
exponential factor for a monomolecular reaction is about 1013 s-1.
28

Table 2. Orders of magnitude of the kinetic parameters for first-order homogeneous reactions
Reaction type Rate coefficient k Activation energy Ea (kJ mol-1)
(s1)

Slow < 103 102 Low < 20

Moderate 102 10 moderate 20 130

Fast > 10 High > 130

In the case of a heterogeneous reaction, the dimension of the rate coefficient for the forward
reaction of (52) is

k
rS
rS (69)
j i
cAi X j cAi
i

i j i
Using normalized surface concentrations, the dimension of the rate coefficient is
2 1
mol mcat s
k 3
2 1
m3mcat s if one gas-phase molecule participates in the reaction (typically
mol m

not more than one gas-phase molecule reacts with the catalyst surface) and k mol mcat
2 1
s if

no gas-phase reactant participates. Presenting the surface concentration in normalized form, we


are automatically including the total concentration of active sites, as t , in the rate coefficient.

5.7. Distinguishing the Kinetic Dependences Based on Patterns


29

Assuming the single reaction AB and the rate of the component change R = k Can, where k is
the kinetic rate coefficient and n is an apparent reaction order, one can distinguish the apparent
order n based on the concentration change in time. The useful concept is a half-time, i.e. time
during which a half of the reactant concentration is consumed.

Concentration Patterns and Fingerprints

(1) The first order reaction r = kCA

A linear dependence ln Ca vs t is observed

ln CA = ln CA0 - kt (70)
A slope is a kinetic coefficient (- k)
Half-life at which CA = 1/2 CA0 is determined as follows:
Half-life = 0.693 / k. It does not depend on the initial concentration
k = 0.693/(Half-life)

(2) The second order reaction, r = kCA2

A linear dependence ln (1/Ca) vs t is observed

(1/ CA) = (1/ CA0) + kt - (71)


A slope is a kinetic coefficient (+k)
Half-life at which CA = 1/2 CA0 is determined as follows:
Half-life = 1/(kCA0) It does depend on the initial concentration. It is inversely
proportional to the initial concentration
k = 1/ [(Half-life) x CA0]

(3) The zero order reaction r = k

A linear dependence Ca vs t is observed

CA = CA0 - kt - (72)
30

A slope is a kinetic coefficient (-k)


Half-life at which CA = 1/2 CA0 is determined as follows:
Half-life = CA0/(2k) It does depend on the initial concentration.
It is proportional to the initial concentration
k = CA0/ 2(Half-life)

(4) The n-order reaction r = kCAn

A linear dependence (1/ CA) (n-1) vs t is observed

(1/ CA) (n-1) = (1/ CA0)(n-1) + (n-1) kt (73)

(5) The first-order reversible reaction A = B, with dependences r+=k+CA and and r- = k-CB,
where r+ and r- are rates of forward and backward rates, respectively; k+ and k-are the
rate coefficients of forward and backwards rates, respectively; CA and CB are
concentrations of substances A and B, respectively.
In the batch reactor, CA + CB = CA0 =1, where CA0 is an initial concentration of
substance A.
A special case of the non-steady-state regime of the reversible reaction is an equilibrium.

At the equilibrium point

k cA,eq k cB,eq (74),

where cA,eq and cB,eq are the equilibrium concentrations of A and B.


At the equilibrium, cA, final = cA, eq ,
Since cA,eq cB,eq cA0 , it that

k k
cA,eq cA0 ; cB,eq cA0 (75)
k k k k
Obviously, cA, eq is not equal 0. Rigorously speaking, the equilibrium concentration
(conversion) is achieved at t .
The incomplete equilibrium conversion is a general characteristic of a reversible reaction,
i.e. the final reactant conversion is not complete.
It can be considered as a general fingerprint for distinguishing the reversible reaction.
For the reversible reaction of the first-order, the linear temporal dependence is observed
31


cA cA,eq cA0 cA,eq exp k k t (76)

Such elegant dependence implies that at any time the concentration distance from the
equilibrium, cA cA cA,eq , can be determined by multiplying the initial concentration distance,
cA 0 cA0 cA,eq , with the exponential term exp( k k t ) , which is similar to that for the
irreversible reaction. The latter is a special case of the reversible reaction; k = 0 and cA,eq = 0
and Eq. (76) reduces to cA cA0 exp kt

In terms of conversion, XA = (cA cA,0) / cA,0 , cA = cA,0 (1-XA) and cA eq = cA,0 (1-XA,eq) where
XA and XA, eq are conversion and conversion at equilibrium condition, respectively, Eq. (76) can
be written as


X A,eq X A X A,eq exp k k t (77)

or


X A X A,eq X A,eq exp( k k t ) X A,eq 1 exp( k k t ) (78)

where XA,eq = k+/( k+ + k) and:

1 k k 1

1 (79)
X A,eq k K eq

with Keq , the equilibrium coefficient (equilibrium constant)

In a batch reactor, the maximum conversion is reached after a certain time and then remains
constant or approximately constant, see Fig. 1. In the case of a reversible first order reaction, the
vicinity of the equilibrium conversion is reached faster than for the corresponding irreversible
reaction (assuming that the forward rate coefficient is the same) because the temporal
exponential change is determined by two rate coefficients, forward and reverse, that work
together. In both cases, a simple transient regime can be observed. The concentration and
conversion dependences reach an equilibrium point without overshoot. Obviously, the
dependence related to the irreversible first order reaction cA cA0 exp kt is a specific case of
Eq. (76) that corresponds to the reversible one at k = 0. .

For both irreversible and reversible reactions of the first order, conversion X does not depend on
the initial reactant concentration cA It is just a function of kinetic coefficients and time. The
half-decay time can be determined as time at which XA = (1/2) XA, eq. For the irreversible
32

reaction case, cA,eq = 0, and XA,eq = 1. In these both cases, the half-decay time is characterized by
the independence on the initial reactant concentration.

Therefore, the independence of the half-decay time on the initial reactant concentration can be
chosen as the fingerprint of the first-order reaction, both irreversible and reversible.

The single reversible reaction with two reactions of the first order presents the simplest example
of the complex reaction which consists of two single reactions, forward and reverse one.

6. Influence of the Temperature on Kinetic Characteristics of the Single Reaction

6.1. Single Irreversible Reaction

As mentioned in Section 5.2, the .Arrhenius dependence of the temperature is

k = ko exp (-Ea/RT), or ln k = ln ko (Ea/RT),

where k is a reaction constant (kinetic coefficient)

If the reaction is monomolecular (first order reaction), k has a unit 1/s;

ko = A is a pre-exponential factor. It has the same unit as k, 1/s.

E is an activation energy, kJ/mol; R is the gas constant (R=8.314 J/mol K); T is the absolute
temperature, K
33

A graph of ln ka vs (1/T) is a straight line.

It is considered that the activation energy and pre-exponential factor do not depend on the
temperature.

At temperatures T1 and T2,

ln k1 = ln ko (Ea/RT1) and ln k2 = ln ko (Ea/RT2), respectively.

Subtracting ln k1 from ln k2, one can obtain:

ln k2 - ln k1 = (Ea/R)[(1/T2) (1/T1)]

Example: HI(g) + CH3(g) CH4(g) + I (g)

This reaction is bimolecular (the second order reaction), and the reaction constant has a unit
L/(mol s)

k1= 7.9 x 10-4 L/ mol s at 2000 C (473.2 K); k2= 0.39 L/ mol s at 3000 C (573.2 K)

(lnk2 / lnk1) = (0.39 L/ mol s) / (7.9 x 10-4 L/mol s) = (-Ea / (8.314 J/mol K) x [( 1/573.2K)-
(1/473.2K)]

Ea = 1.4 x 105 J/mol (1 kJ / 103J) = 1.4 x 102 kJ/mol

To determine the pre-exponential factor k0 = A, we have to use one of k values, k1 or k2

Using k2 and T2, 0.39 L/mol s = ko x exp [( - 1.4 x 105 J/mol] / ((8.314 J/mol K) x 573.2 K)]

k0 = (0.39 L/mol s) / (1.74 x 10-13) = 2.2 x 1012 L/mol s

As mentioned, the kinetic coefficient, k, and the pre-exponential factor, k0= A, have the same
units, i.e. L/(mol s)
34

6.2. Single Reversible Reaction.

Fig. 2 Energetic Diagram: Dependence of the Potential Energy on the Reaction Progress

Ea, f and Ea, b are activation energies of the forward and reverse reactions;

As an example, the following reaction is chosen (Fig.3)

NO2 (g) + CO (g) NO (g) + CO (g)

g means gas.

Ea, f = 132 kJ/ mol; Ea, b = 358 kJ/mol


35

Heat of reaction H = Ea, f - Ea, b = 132 kJ/ mol - 358 KJ/mol = - 226 kJ/mol

In this case, H 0. It means that heat is released, and the reaction is exothermic.

If H heat of reaction has to be supplied, and the reaction is endothermic.

For reaction mentioned NO2 (g) + CO (g) NO (g) + CO (g), the forward reaction is
exothermic. Hence, the corresponding backward (reverse) reaction

NO (g) + CO (g) NO2 (g) + CO (g) is endothermic.


36

6.3. Influence of the Temperature on the Equilibrium Composition of the First Order Reaction.

As mentioned, the equilibrium is a characteristic of the reversible process.

Under equilibrium conditions, the rate of forward reaction (Rf) equals to the rate of the backward
(reverse) reaction (Rb), Rf = Rb

The simplest example is an isomerization reaction A B, where A and B are cis- and trans-
isomers of butane, C4H10

Rf = kf CA and Rb = kb CB-,

where CA , CB- are concentrations of the reactant A and product B, respectively,

kf , kb- are kinetic coefficients (reaction constants) of the forward and reverse reactions,
respectively.

Under equilibrium conditions:

kf CA = kb CB, see Eqn, 74

and, hence,

(kf / kb ) = (CB / CA ),

Keq = (kf / kb) is the equilibrium constant.

Q = (CB / CA) is a quotient of the reaction.

Properties of the Equilibrium Constant

(1) It is a function of the only temperature, not mixture composition


(2) Its value is independent on how the equilibrium has been reached
(3) It has to be distinguished two values of Keq , Kc and K p

Kc is calculated as a function of concentrations. Typically, Kc is calculated as a


dimensionless value since all concentrations are taken as dimensionless. Every
concentration is taken as a ratio of its absolute value divided by the standard
concentration (1 mol).
K p is calculated for gases as a function of partial pressures.
In difference from K c , K p has a dimension which depend on the type of reaction

Kc = K p (1/RT) n ,
37

and

K p = Kc (RT) n,
where n is a difference between the number of moles of products and number of moles
of reactants, respectively

Equilibrium constant can be determined using vant Hoffs equation

ln (K eq.2 / K eq.1) = (- H/R) [(1/T2) (1/T1)],

where K eq.1 are K eq.2 equilibrium constants at temperatures T1 and T2 , respectively.

For T2 T1 (temperature rise), if H 0 (exothermic case), K eq.2 K eq.1

If H 0 (endothermic case), K eq.2 K eq.1

6.4. Equilibrium Constant and Quotient

Obviously, Keq = Q under equilibrium conditions.

kf = Af exp (- E+/RT); k b = Ab exp (- E-/RT)

Example

What is (K eq.2 / K eq.1), if H = 1.00 x 102 kJ/mol x (103 J/ 1kJ) and the temperature increase is
from 20o C (T1 = 293.15 K) to 30oC (303.15 K)

ln (Keq.2/ Keq.1) = (- 1.00 x 105 J/mol) [(1/303.15 K) (1/293.15 K)] / 8.314 J/mol K = + 1.353

(Keq.2/ Keq.1) = 3.87 ( Keq.2 Keq.1 )

As mentioned, under equilibrium conditions the equilibrium constant and quotient are equal

K eq. = Q

For isomerization reaction A B,

Keq = (k+/k-) = (Cb,eq /Ca,eq) = Q

For dimerization reaction 2 NO2 2 N2O4 under equilibrium conditions


38

Keq = (k+/k-) = (CN2O4,eq /C2NO2,eq) = Q

Under non-equilibrium conditions,

K eq Q

If Q Keq, the reverse (backward) reaction dominates. In this case, the chemical reaction is
shifted to the left side (side of reactants).

If Q Keq, the forward reaction dominates. In this case, the chemical reaction is shifted to the
right side (side of products)

6.4. Calculations of the Equlibrium Constant and Reaction Quotient.

There are three typical problems in this area:

(1) To calculate Kc using equilibrium concentrations;


(2) To calculate equilibrium concentrations using Kc;
(3) To calculate a quotient of the reaction,Q, using the concentrations.
Then, to determine a direction of the reaction comparing Q with Kc

Equilibrium and Free Gibbs Energy

K = exp (-Grxn /RT),


where Grxn is a standard value of the free Gibbs energy.
Typical questions are
(1) How to use tables for calculating Grxn?
(2) What are standard values of Grxn ?
(3) How to calculate Grxn for the given reaction?
(4) Is the reaction spontaneous or not?
For the answering the last question, we have to calculate Grxn at the given
conditions. If Grxn is negative, the reaction is occurred to be spontaneous.
If Grxn is positive , the reaction is considered to be non-spontaneous

6.5. Equilibrium and Optimum Conversion (Comparison)*


* G. Yablonsky, and A. Ray, Equilibrium and Optimum: How to Kill Two Birds with One Stone,
International Journal of Chemical Reactor Engineering, 6(2008 )A36

(A) Equilibrium Conversion

For the reversible reaction A B, assuming kf CA,eq = kCb,eq.,. and


39

CA = CA,0 (1-X); CB = CA0 X, where CA0 and CB0 are initial concentrations of substances A and
B, respectively; X is conversion; CA + CB = CA0 + CB0 =1
X = (CA0 - CA) / CA
one can obtain kf CA,0eq (1-Xeq)= kbCA,0 Xeq and Xeq = (kf /kb ) / [1 +(kf /kb )];
Xeq = Keq /(1 + K eq)
For the exothermic reaction, Keq is decreased with the temperature rise. Hence, the equilibrium
conversion is decreased as well.
For endothermic reaction, Keq is increased with the temperature rise. Hence, the equilibrium
conversion is increased as well.

(B) . Optimal Conversion

Optimal conversion is presented as a characteristic of performance of the open chemical reactor


under non-equilibrium conditions
kf CA,eq kCb,eq
In this case, R = kf CA,eq - kCB,eq 0,
where kinetic coefficients are determined by Arrhenius dependences
kf = Af exp (- E+/RT); k b = Ab exp (- E-/RT)
and CA = CA,0 (1-X); CB = CA0 X

Optimal conversion is determined as follows

(R/T ) = 0

and

Af exp (- E+/RT) (E+/RT2) CA,0 (1-Xopt) - Ab exp (- E-/RT)( E-/RT2 ) CA0 Xopt = 0

Hence, Keq (1-Xopt) = Xopt, where = E+/ E-

and Xopt = (Keq ) / (1 + Keq )

For the exothermic and endothermic reaction = (E+/ E-) is less than 1
and more than 1, respectively.
The detailed analysis shows that the optimal maximum regime is realized only for the
exothermic reaction
As for the endothermic regime, the optimal maximum regime is not realized. Conversion is
increased with the temperature rise. Conversion will be limited just by the technological
requirements, e.g. catalyst stability, safety recommendations etc.
Comparison between the equilibrium and optimal conversion
40

Xeq = Keq / (1 + Keq ) and Xopt = (Keq ) / (1 + Keq )


shows that for the exothermic reaction ( is less than 1) optimal conversion is always less than
equilibrium conversion
X opt X eq
For the endothermic reversible reaction, the optimal temperature policy is: to keep the
temperature as high as possible under the requirements that are governed by the technological
limitations (safety requirements, catalyst stability etc)

7. Distinguishing the Reactions and Kinetic Dependences

7.1 Distinguishing Single and Multiple Reactions

In a single reaction (elementary or complex), the ratios between rates of changes of components
are presented by the integer (whole) numbers. These numbers which are not necessary 0, 1, 2,
and 3 correspond to the stoichiometric coefficients of the reaction. Stoichiometric equations for
real complex reactions are free from the limitations that are set on the stoichiometric
coefficients of elementary reactions (1, 2 and rarely 3). For example, in the C2H4O oxidation
reaction, rates of changes of components C2H4O, O2 , CO2 and H2O can be presented as ratios
of integer numbers 2, 5, 4 and 4, respectively. For this reaction, the stoichiometric equation is

2 C2H4O + 5 O2
4 CO2 + 4 H2O (80)

We can consider this reaction as the single one, but not the elementary one.

Moreover, this ratio has to be constant or approximately constant

If these ratios of component rates do not correspond to the stoichiometric relationships, the
reaction cannot be considered as the single one. It is a case of multiple reactions.

7.2. Distinguishing the Power Law and Non-Power Law Dependences.

A goal of this procedure is to verify a validity of the power law expression say
r k cA11 cA22 k cAii
41

This test will be a basis of further steps: (1) construction of the model for reactor design or /and
(2) development of the mechanistic model that corresponds to the detailed mechanism of the
complex reaction. Assuming the reaction is irreversible, and the reaction rate is governed by the
power law expression

r k cA11 cA22 k cAii

values of powers can be determined varying the initial composition of reactive mixture.

Distinguishing power low dependences is occurred using patterns which have been described in
Section 5.7. In this procedure, there are no restrictions for power values. Typically, they are
equal 0, 1, 2, 3 (rarely). They can be fractional, positive or negative. The obtained power law
dependence is considered as the empirical or rather to say semi-empirical expression.

However simple power law dependences are invalid to describe the experimental data in which
power coefficients are changed depending on the concentration and temperatures domain
studied.

The typical non-power law kinetic dependence is the equation by Langmuir and Hinshelwood in
the 1920s-1930s (LH-equation) for describing catalytic oxidation reaction, in particular CO and
hydrogen oxidation over metals of the platinum group

k cini
r i
(81)
1 Ki cini
i

The Langmuir-Hinshelwood equation reflected complexity of the catalytic reaction: (i) a


catalytic process involves competition between the components of the reaction mixture for sites
on the catalyst surface, and (ii) very often adsorption and desorption rates are potentially high in
comparison with other steps of chemical transformation on the catalyst surface.

Then, , Hougen and Watson in the 1940s proposed a similar semi-empirical rate equation, the
HW equation, for a reversible complex catalytic reaction, considering it the basis of reactor
design. For instance, for the dehydrogenation of cyclohexane this equation has the form
42

k cC6H12 k cC6H10 cH2


r (82)
( K1cC6H10 K 2cH 2 K3cC 6H12 )m

Hougen and Watson also presented the denominator in the form

(1 KA aA KR aR KI aI )n , (83)

where aA , aR , aI , are thermodynamic activities of reactants and products and KA, KR, KI,

are adsorption equilibrium coefficients of reactants and products.

Equations (81) - (82) are called Langmuir-Hinshelwood-Hougen-Watson (LHHW)-equations


indicating their similarity: they belong to the class of numerator-divided-by denominator-
kinetic equations. Within the HW approach, a rate-determining step is typically assumed. In
many cases the adsorption/desorption steps are considered to be quasi-equilibrated
In HW equations, the numerator of the rate expression is considered to correspond to the net
reaction.

For the LHHW-equations there are two physico-chemical requirements for LHHW-equations:

(a) All coefficients of these equations have to be positive


(b) If the numerator of the LHHW-equation has two terms which correspond to forward and
reverse reactions respectively, with coefficients k+ and k-, the relationship
(k+ / k- ) = K eq,n
has to be fulfilled, where K eq,n is the equilibrium constant for the net reaction. It is the
absolutely necessary thermodynamic requirement.

The special power law kinetic equation is the well known Temkin-Pyzhev equation for
catalytic reaction of ammonia synthesis over the iron catalyst
N2 + 3 H2 = 2 NH3 (84)
43

R = f {[k+ (PN2 PH21.5)/PNH3] - [k- (PNH3/ PH21.5)]}, (85)


where PN2 , PH2 ,PNH3 are partial pressures of nitrogen, hydrogen and ammonia,
respectively; f is the factor of the catalyst activity; k+ and k- are kinetic coefficients for
the forward and reverse reaction of ammonia synthesis, respectively;
(k+ / k- ) = Keq , where Keq is the equilibrium constant of reaction (84)
It is shown that the Temkin-Pyzhev equation is a particular case of the LHHW-equation
(see http://www.ncsu.edu/checs/amcs/am_prob2.htm)

Eqn (85) can be easily transformed into the LHHW equation

R = f K3 {[k+ (PN2 PH21.5)] - [k- (PNH32/ PH21.5]} x 1/ (1 + K3 PNH3), (86)


where K3 is a kinetic coefficient which reflects ammonia adsorption

7.3. Kinetic Decoding the Elementary Reactions from Complex Reaction.

Decoding the complex kinetic dependence in terms of elementary reaction is a quite complicated issue.
This procedure is based on the concepts of the mass-action-law and Arrhenius temperature dependence.
For the elementary reactions, mono-, bi-, and termolecular, the allowed stoichiometric coefficients are
only integer (whole ) numbers 1, 2 and 3. The last coefficient is occurred quite rarely. Moreover, the
kinetic dependence of the elementary reaction has to be a power law with the same stoichoichiometric
coefficients 1, 2 and 3 (rarely), and the sum of these coefficients must be not be larger than three.
Also the temperature dependence of the kinetic coefficient has to be of Arrhenius type, i.e. exponential.
The parameters of interest (kinetic orders and activation energies) have to be constant or approximately
constant in the studied domain of temperatures and concentrations.

Regarding these strong requirements, there are three necessary conditions for selecting the elementary
reactions

A. Small Stoichiometric Coefficients: 1,2 and rarely 3


Ratios of component rates have to be integer numbers 1, or 2, or 3, not larger related to these
small stoichiometric coefficients related to the number of reacting molecules. Zero kinetic order
is a fingerprint of the complex, not elementary reaction.
44

B. .Sum of Kinetic Orders: not larger than three.


In fact, it is also the physical requirement. The sum of these coefficients must also not be
larger than three because a probability of the collision more three molecules is
insignificant
C. Arrhenius Dependence of Kinetic Coefficients
The temperature dependence of kinetic coefficients has to be exponential, see Eqns (40), (41).

As already mentioned in the Section 5.1, most chemical reactions are complex, i.e. consist of many
elementary reactions. Typically, kinetic dependences of chemical reactions do not meet all
requirements A, B and C. Hence, the chemical reaction via one elementary reaction is rather a
unique exception. As the result, for the single reaction the observed kinetic dependence is
typically not the mass-action law expression.

E.g., the Water-Gas-Shift (WGS) catalytic reaction

H2O + CO = H2 +CO2

is occurred via the two-step mechanism:

(1) H2O + Z =ZO + H2; (2) CO + ZO = Z + CO2, (87)

where Z is a catalyst (empty site), ZO is an oxidized site. It is the simplest two-step catalytic
mechanism.

We can start developing the kinetic description of the WGS-reaction based on the simple kinetic
dependences (mass- action-law dependences), i.e. equation (38) and (39), and the net-rate of this
reaction will be presented as

R = k+PH2OPCO k-PH2PCO2 , (88)


where PH2O , PCO, PH2 and PCO2 are partial pressures of substances H2O, CO, H2 and
CO2, respectively.

This dependence can be used as a primary kinetic model in some domain of parameters
(concentration and temperatures) which may be wide enough. In kinetic modeling, it is a
reasonable simplification, which gives a ground for decoding the complex dependence.
45

Rigorously, the theoretical kinetic dependences for describing the WGS-reaction kinetics in a
wide experimental data domain are much more complicated, e.g. Langmuir-Hinshelwood-
Hougen-Watson (LHHW) dependences, because this reaction is not elementary. However in
special parametric domains of interest, these complex dependences can be significantly
simplified up to the dependence (88) and even more simple power law kinetic dependences.

For example, the reaction rate for the WGS-reaction can be

R = k Pco , or R = k PH2O, or even R = k, (89)

Such simplification can be interpreted as follows:

if the WGS-reaction in some domain is considered to be irreversible, and its mechanism consists
of two irreversible steps (87), the complex reaction rate is expressed by the relationship

R= (k1PH2O ) (k2PCO) / (k1 PH2O + k2PCO) (90)

Under conditions, that k1PH2O is much larger k2PCO,

R= k2PCO , i.e. the kinetic characteristic of the second elementary reaction.

Symmetrically, under conditions that k2PCO is much larger k1PH2O,

R= k1PH2O, i.e. the kinetic characteristic of the first elementary reaction.

It means that in some experimental domains the complex kinetic dependence being simplified
reveals the details of the chemical mechanism, i.e. its elementary reactions.

Summing up:

Certainly the chemical reaction is complex, not elementary, if kinetic orders are negative,
fractional (positive and negative) and even zero. Also the reaction is not necessarily elementary
if only one of the above mentioned statements is true. For example, many reactions in which
only one, two or three components participate are not elementary. Furthermore, in some cases
the kinetic law of a complex reaction may be approximated by the simple kinetic mass-action-
law power law with powers 1, 2 or 3; however it is just an approximation. In reality, the reaction
studied is occurred through the complex detailed mechanism with many steps
46

It has to be stressed that conditions A, B and C are only necessary, not sufficient conditions for
distinguishing the elementary reactions from the complex ones. Also the numerical values of the
obtained parameters, in particular pre-exponential factors and activation energies, have to be
analyzed to justify them from the physico-chemical point of view. E.g., the typical pre-
exponential factor for the first order reaction is 1013, 1/s etc.

A result of our analysis is the following one:

(1) Rigorously, the observed kinetic data could not be described based on the net
stoichiometric reaction (global reaction) as an elementary reaction.
(2) Rigorously, the observed kinetic data could not be described based on the mass-
action-law (MAL) equation related to the net stoichiometric equation

(3) However in some wide parametric domains, the kinetic data can be described by the
simple power law dependences. Being simplified in some experimental domains, the
complex kinetic dependence reveals the details of the chemical mechanism, i.e. its
elementary reactions.

7.4. Four Types of Kinetic Scenarios


Generally, four types of kinetic scenarios can be distinguished
A. Single reactions with simple power law kinetic dependences
B. Single reactions with the complex, non-power -law kinetic dependences
C. Multiple reactions with simple power law dependences
D. Multiple reactions with non- power -law dependence.
We discussed types A and B previously. Since 1960s and up to now the main challenge is
the development of kinetic descriptions for multiple reactions. First, it has to be assumed
a network of reactions based on the experimental information and known fingerprints.

In decoding chemical complexity, there is a typical problem of parallel versus


consecutive reactions, see Fig. 4.
47

k1 B k1 k2
A B C
A
k2 C

(a) (b)

Fig. 4 (a) Parallel reactions; (b) Consecutive reactions.

The question is which mechanism is the right one mechanism for the formation of the desired
product B. For example, in any hydrocarbon (say substance A) oxidation process we are
interested in knowing whether products CO (substance B) and CO2 (substance C) are produced
in parallel reactions, or CO is produced first, and, then, CO2 second.

Fig. 5 Qualitative temporal kinetic dependences in a batch reactor for the parallel reactions.
48

Fig. 6 Qualitative temporal kinetic dependences in a batch reactor for the consecutive
reactions.

The difference between the parallel and the consecutive reactions is very clear (Fig. 5 and Fig. 6)
For the consecutive reactions, the concentration of B goes through a maximum. This is a
characteristic of the consecutive reactions..

In both cases, the material balance in the batch reactor has to be fulfilled for each of the
participating components:

c c
i
i
i
i0 with i A, B, C (91)

Assuming only A is present initially, the initial conditions are

cA cA0 ; cB 0; cC 0 (92)
Lets assume the simplest power low dependence for all reactions, i.e. the reaction rate for any
reaction will be the first order expression (linear dependence)

The model for the parallel reactions consists of the following equations:
49

dcA
k1cA k2 cA k1 k2 cA
dt
dcB
k1cA (93)
dt
dcC
k2 cA
dt

where k1 and k2 are the rate coefficients of the corresponding reactions.

The solution to this model is

cA cA0 exp k1 k2 t
k1
cB cA0 1 exp k1 k2 t (94)
k1 k2
k2
cC cA0 1 exp k1 k2 t
k1 k2
The kinetic dependences are quite simple; the concentration of A decreases
exponentially, while the concentrations of B and C increase exponentially, see Fig. 5.

The model for the consecutive reactions consists of the following equations:

dcA
k1cA
dt
dcB
k1cA k2 cB (95)
dt
dcC
k2 cB
dt

where k1 and k2 are the rate coefficients of the corresponding reactions.

The solution to this model is


50

cA cA0 exp k1t


k1
cB cA0 exp k1t exp k2t (96)
k2 k1
k2 k1
cC cA0 1 exp k1t exp k2t
k2 k1 k2 k1

and, again, summation of the three concentrations yields cA0 . Fig. 6 shows the kinetic
dependences for the consecutive reactions.

Solutions (94) and (96) are fingerprints for parallel and consecutive reactions, respectively. They
can applied both to the batch reactor (BR), and plug-flow reactor (PFR).

It is possible that experimental data correspond to the visual fingerprint say for the
consecutive reactions (Fig. 4b and Fig. 6), but at the same time they dont fit to the
analytical dependence (96). In this case, it has to be concluded that the kinetic law of
reaction has to be changed. It has to be assumed a new modified power law or even a
non-power kinetic law of LHHW-type.
In the design of the multiple reaction networks, the reasonable recommendation can be
done. The development of such networks has to be based on experimental information on
the chemical composition change, and the primary number of reactions has to be
determined (equal) to the number of chemical concentrations measured independently.
Rates of assumed reactions can be found directly from the rates of change of measured
substances. One can define such a set of reactions as a minimal one.
It is important to stress that the real set of multiple reactions is not necessarily a minimal
set; nature does not necessarily only meet the minimum requirements. However, the
minimal set of reactions can be used as a starting point for revealing the real set.
Also the desired kinetic model has to be characterized by the optimal complexity, i.e.
to be not too simple, not too complex.
51

NOTATION

Symbol Description Units

Ai Reactant -

Bi Product -

cAi concentration of reactant i mol m-3

cBi concentration of product i mol m-3

Ea activation energy J mol-1

rH reaction enthalpy J mol-1

Keq equilibrium constant -

k reaction rate coefficient (mol m-3)(1-n)s-1

k0 pre-exponential factor same as k

nAi number of moles of reactant i Mol

nBi number of moles of product i Mol

Rg universal gas constant J mol-1 K-1

Ri net rate of production of component i mol m-3s-1

R reaction rate for a homogeneous reaction mol m-3s-1

rS reaction rate per unit surface area of catalyst 2 1


mol mcat s

rV reaction rate per unit volume of catalyst 3 1


mol mcat s

rW reaction rate per unit mass of catalyst 1 1


mol kg cat s

Scat catalyst surface area 2


mcat
52

T Temperature K

V reaction volume m3

Vcat volume of catalyst in the reactor m3cat

Vf volume of fluid in the reactor m3

Wcat mass of catalyst in the reactor kgcat

W reaction weight 2 1
mol mcat s

Xj, Yj surface intermediates -

Greek symbols

i absolute value of stoichiometric coefficient of reactant Ai -

j absolute value of stoichiometric coefficient of intermediate Xj -

i stoichiometric coefficient of product Bi -

j stoichiometric coefficient of intermediate Yj -

X j concentration of surface intermediate Xj 2


mol mcat

t total concentration of surface intermediates 2


mol mcat

X j normalized concentration of surface intermediate Xj -

i stoichiometric coefficient of component i -

cat catalyst density 3


kgcat mcat
53

Subscripts
Cat Catalyst

S Step

Site with respect to one site

Superscripts
+ of forward reaction

of reverse reaction

i partial order of reaction in reactant Ai

j partial order of reaction in intermediate Xj

i partial order of reaction in product Bi

j partial order of reaction in intermediate Yj

N order of reaction

You might also like