Hemical Ngineering Inetics: Based On CHEM - ENG 408 at Northwestern University
Hemical Ngineering Inetics: Based On CHEM - ENG 408 at Northwestern University
Hemical Ngineering Inetics: Based On CHEM - ENG 408 at Northwestern University
TABLE OF CONTENTS
1 Reactor Archetypes ............................................................................................................... 3
1.1 The Mass Balance ......................................................................................................................... 3
1.2 Batch Reactor ................................................................................................................................ 3
1.3 Continuous-Stirred Tank Reactor ................................................................................................. 4
1.4 Plug-Flow Reactor ........................................................................................................................ 4
2 Arrhenius Expression ............................................................................................................ 5
2.1 Arrhenius Equation ....................................................................................................................... 5
2.2 Van’t Hoff Equation ..................................................................................................................... 5
3 Mass-Action Kinetics............................................................................................................. 7
3.1.1 Overview .................................................................................................................................................7
3.1.2 First Order Kinetics .................................................................................................................................7
3.1.3 N-th Order Kinetics .................................................................................................................................7
3.1.4 Reversible Reactions................................................................................................................................8
3.1.5 Non-Integer Rate Laws ............................................................................................................................9
3.2 Determining the Rate Law and Rate Constants ............................................................................ 9
3.2.1 Differential Method .................................................................................................................................9
3.2.2 Integral Method ..................................................................................................................................... 10
3.2.3 Regression Method ................................................................................................................................ 10
3.2.4 Working with Pseudo Orders ................................................................................................................. 10
4 Working with Multiple Elementary Steps ........................................................................ 11
4.1 General Approach ....................................................................................................................... 11
4.2 Synthesis of HBr ......................................................................................................................... 11
4.2.1 Elementary Steps ................................................................................................................................... 11
4.2.2 Rate Expressions .................................................................................................................................... 11
4.2.3 The Pseudo-Steady State Hypothesis..................................................................................................... 12
4.2.4 Bond Dissociation Energies ................................................................................................................... 12
4.3 Complex Reactions: Cracking .................................................................................................... 12
4.3.1 Terminology .......................................................................................................................................... 12
4.3.2 Mechanism and Mass-Action Kinetics .................................................................................................. 13
4.3.3 Simplifications Based on Concentrations .............................................................................................. 14
4.3.4 Simplifications Based on Statistical Termination .................................................................................. 14
4.3.5 Simplifications Based on the Long-Chain Approximation .................................................................... 14
4.3.6 Ethane Cracking Rate Law .................................................................................................................... 14
4.3.7 Determining Observed Activation Energies .......................................................................................... 15
4.3.8 General Overview of Simplification Process ......................................................................................... 15
4.3.9 Complex Reactions: Additives .............................................................................................................. 16
4.4 Complex Reactions: Radical Chain Autoxidation ...................................................................... 16
5 Consequences of Chemical Equilibria ............................................................................... 18
5.1 Relationships Between Thermodynamics and Equilibrium ........................................................ 18
5.1.1 Temperature-Dependence of Thermodynamic Quantities ..................................................................... 18
5.2 The Equilibrium Constant ........................................................................................................... 18
5.2.1 Activity Equilibrium Constant ............................................................................................................... 18
5.2.2 Activities for Gases ................................................................................................................................ 19
5.2.3 Activities for Liquids ............................................................................................................................. 19
5.2.4 Relationship Between Various Equilibrium Constants .......................................................................... 20
5.3 Enzyme Kinetics ......................................................................................................................... 21
5.3.1 Derivation of the Michaelis-Menten Equation ....................................................................................... 21
5.3.2 Plotting Michaelis-Menten Data ............................................................................................................ 22
5.3.3 Reversible Product Binding ................................................................................................................... 23
5.3.4 Competitive Inhibition ........................................................................................................................... 24
5.3.5 Non-Competitive Inhibition ................................................................................................................... 24
6 Reaction Networks............................................................................................................... 25
REACTOR ARCHETYPES | 2
1 REACTOR ARCHETYPES
1.1 THE MASS BALANCE
The key equation governing processes on the reactor level is the mass balance. In order to inherently account
for the proper stoichiometry, this is most typically written as a mole balance. The general mole balance for
a species 𝑖 is given as
𝑑𝑁𝑖
𝐹𝑖0 − 𝐹𝑖 + 𝐺𝑖 =
𝑑𝑡
where 𝐹𝑖0 is the input molar flow rate, 𝐹𝑖 is the output molar flow rate, 𝐺𝑖 is the generation term, and the
differential term is the accumulation.
If the system variables are uniform throughout the system volume, then
𝐺𝑖 = 𝑟𝑖 𝑉
where 𝑟𝑖 is the reaction rate of species 𝑖 and 𝑉 is the system volume. More generally speaking, if 𝑟𝑖 changes
with position in the system volume, then
𝐺𝑖 = ∫ 𝑟𝑖 𝑑𝑉
1 𝑑𝑁𝑖 𝑑𝐶𝑖
𝑟𝑖 = =
𝑉 𝑑𝑡 𝑑𝑡
where 𝐶𝑖 is the concentration of species 𝑖 and the expression 𝑁𝑖 = 𝐶𝑖 𝑉 was utilized.
Occasionally, batch reactors can be operated at a constant pressure but with a system volume that changes
as a function of time. In this special case,
1 𝑑𝑁𝑖 1 𝑑 𝑑𝐶𝑖 𝐶𝑖 𝑑𝑉
𝑟𝑖 = = (𝐶𝑖 𝑉(𝑡)) = +
𝑉(𝑡) 𝑑𝑡 𝑉(𝑡) 𝑑𝑡 𝑑𝑡 𝑉(𝑡) 𝑑𝑡
since the system is in steady state. Solving for the rate, dividing by Δ𝑉, and letting it approach zero (while
applying the definition of the derivative) yields
𝑑𝐹𝑖
𝑟𝑖 =
𝑑𝑉
Once again substituting in for 𝜏 ≡ 𝑉/𝑣̇ and 𝐹𝑖 = 𝐶𝑖 𝑣̇ (assuming steady state conditions such that 𝑣̇ is
constant) yields
𝑑𝐶𝑖
𝑟𝑖 =
𝑑𝜏
Note that the design equation can be written in terms of the length of the reactor, 𝑧, and the cross-sectional
area, 𝐴𝑐 , if 𝑉 ≡ 𝐴𝑐 𝑧. Occasionally, with PFRs, the superficial velocity will be referred to, which is simply
𝑢 ≡ 𝑣̇ /𝐴𝑐 .
2 ARRHENIUS EXPRESSION
2.1 ARRHENIUS EQUATION
The main assumption behind the Arrhenius expression is that 𝑟𝑖 = 𝑓(𝑇) ⋅ 𝑓(𝐶𝑖 ). This is an approximation,
but it works quite well. The rate coefficient is the term that is a function of temperature but may also depend
on things like catalyst, solvent, and such. Empirically, the Arrhenius expression states that
𝐸
𝑘 = 𝑘0 𝑒 −𝑅𝑇
where 𝑘 is the rate coefficient, 𝑘0 is the pre-exponential factor, 𝐸 is the activation energy, 𝑇 is the
temperature, and 𝑅 is the ideal gas constant. By linearizing the equation, one finds that
𝐸 1
ln(𝑘) = − ( ) + ln(𝐴)
𝑅 𝑇
such that plotting ln(𝑘) vs. 1/𝑇 should yield a straight line of slope −𝐸/𝑅 and 𝑦-intercept of ln(𝐴).
To find the ratio of two rate coefficients and two temperatures,
𝑘2 𝐸 1 1
= exp (− ( − ))
𝑘1 𝑅 𝑇2 𝑇1
3 MASS-ACTION KINETICS
3.1.1 OVERVIEW
We assume that rates can be described by
Rate = 𝑓(𝑇) ⋅ 𝑔(𝐶𝑖 )
where 𝑓(𝑇) is the rate constant from Arrhenius’ law and 𝑔(𝐶𝑖 ) is the rate law. Empirically, we say that
𝛽 𝛾
𝑔(𝐶𝑖 ) = 𝐶𝐴𝛼 𝐶𝐵 𝐶𝐶 …
which is typical rate law kinetics. In this expression, if 𝛼, 𝛽, 𝛾, … are the stoichiometric coefficients of a
reaction, it may be an elementary step. Typically, we find that for elementary steps ∑ 𝛼𝛽𝛾 … < 3.
which becomes
𝐶𝐴
− ln ( ) = 𝑘𝑡
𝐶𝐴0
and thereby
𝐶𝐴 = 𝐶𝐴0 𝑒 −𝑘𝑡
and a plot of ln(𝐶𝐴 ) vs. 𝑡 should be linear for first order kinetics. The half-life can be given by
ln(2)
𝑡1 =
2 𝑘
𝐶𝐴 𝑡
1
− ∫ 𝑛 𝑑𝐶𝐴 = 𝑘 ∫ 𝑑𝑡
𝐶𝐴
𝐶𝐴0 0
which becomes
𝐶𝐴−𝑛+1 −𝑛+1
𝐶𝐴0
− = 𝑘𝑡
−𝑛 + 1 −𝑛 + 1
For 𝑛 ≠ 1 and thereby
1 1
= + (𝑛 − 1)𝑘𝑡
𝐶𝐴𝑛−1 𝑛−1
𝐶𝐴0
and a plot of 1/𝐶𝐴𝑛−1 vs. 𝑡 should be linear. The half-life can be given by
2𝑛−1 − 1
𝑡1 = 𝑛−1
2 (𝑛 − 1)𝑘𝐶𝐴0
The units of the rate constant for an 𝑛th order reaction can be given by 𝑘[=]concentration1−𝑛 s −1.
𝑘𝑟
𝐶𝐴,𝑒𝑞 = 𝐶
𝑘𝑓 + 𝑘𝑟 𝐴0
Therefore,
𝑑𝐶𝐴
− = (𝑘𝑓 + 𝑘𝑟 )(𝐶𝐴 − 𝐶𝐴,𝑒𝑞 )
𝑑𝑡
which can be integrated to
𝐶𝐴 𝑡
𝑑𝐶𝐴
− ∫ = (𝑘𝑓 + 𝑘𝑟 ) ∫ 𝑑𝑡
𝐶𝐴 − 𝐶𝐴,𝑒𝑞
𝐶𝐴0 0
which is equal to
𝐶𝐴0 − 𝐶𝐴,𝑒𝑞
ln ( ) = (𝑘𝑓 + 𝑘𝑟 )𝑡
𝐶𝐴 − 𝐶𝐴,𝑒𝑞
The second method is the polynomial method, wherein a polynomial of suitable order is fitted to the data.
The derivative can then be evaluated by differentiating the polynomial expression (typically, the simplest
monotonic function is the best choice).
The third method is the finite difference method, wherein we can approximation the derivative via a
numerical finite difference equation. This, however, requires a sufficient data size to do accurately.
to solve for 𝑛. This can be done through a non-linear least-squares analysis (a typical fitting toolbox would
suffice).
Therefore, based on the BDEs listed, we can assume that 𝑘1 ≫ 𝑘4 due to the low BDE of Br2 compared to
H2. This BDE analysis also means that [Br ∙] ≫ [H ∙]. As such, we will remove the 𝑘4 term. We also note
that since there is so little [H ∙] in the system, 𝑟−4 ≈ 0. We will also consider reaction 3 only in the forward
direction because of the relatively high BDE of HBr. Applying the PSSH and including these assumptions
results in the following solution
1
𝑘[H2 ][Br2 ]2
𝑟=
[HBr]
𝑘′ +
[Br2 ]
At low conversion (i.e. [HBr] ≈ 0), we see that the equation becomes first order in H2 and half order in Br2.
Looking at limiting cases like this is helpful in accurately describing the kinetics.
Hydrogen abstraction takes the general form of X· + H-Y → X-H + Y·. The radical that is produced from
a hydrogen abstraction process is called a 𝜇 species. It is literally the abstraction of a hydrogen from an
otherwise stable substrate.
𝛽-scission takes the following general form:
𝛽1 𝜇
𝛽1 𝑘𝑡 ′ 𝑘𝑡
𝜇 𝑘𝑡 𝑘𝑡 ′′
If we assume statistical termination, then the 𝑘𝑡 reaction should occur twice as likely as the 𝑘𝑡 ′ and 𝑘𝑡 ′′
reaction. As such, if we let 𝜔 be the rate constant for termination, we can state that
1
𝜔 = 𝑘𝑡′ = 𝑘𝑡′′ = 𝑘𝑡
2
4.3.5 SIMPLIFICATIONS BASED ON THE LONG-CHAIN APPROXIMATION
The long-chain approximation can be made when the rate of propagation is significantly faster than the rate
of initiation and/or termination.
In order to better understand what is going on in this mechanism, a diagram can be drawn like the one
below. We see that there is a radical ratcheting cycle between the ethyl radical and hydrogen radical. This
controls most of the kinetics of the process (hence the presence of 𝑘1 and 𝑘11).
This shows us that the overall activation energy for this reaction is
1 1
𝐸𝑜𝑏𝑠 = 𝐸11 + 𝐸𝛼 − 𝐸𝜔
2 2
This approach can be used anytime rate coefficients are multiplied and/or divided together. If rate
coefficients are added or subtracted, the only way to find an observed activation energy is to break it into
different regimes where the addition/subtraction disappears.
From this, we can see that 𝑘4 is likely a small quantity, and so the long-chain approximation probably isn’t
best to use here.
If we want to find the selectivity of ROH versus R=O, we note that ROH is produced in both reaction 1 and
the termination step while R=O only is produced in the termination step. The selectivity is then
𝑑[ROH]
𝑑𝑡 step 1 + step 𝑡 𝑘1 [RO ∙][RH] + 𝑘𝑡 [ROO ∙]2 𝑘𝑑 [ROOH]
= = = +1=2
𝑑[R = O] step t 𝑘𝑡 [ROO ∙]2 𝑘𝑡 [ROO ∙]2
𝑑𝑡
CONSEQUENCES OF CHEMICAL EQUILIBRIA | 18
∘
The Δ𝐻rxn,𝑇𝑟𝑒𝑓
term refers to the standard enthalpy of reaction at a tabulated reference temperature. This is
typically known from the heats of formation of each species at the same reference temperature (Δ𝐻𝑓∘𝑖,𝑇𝑟𝑒𝑓 ).
The relationship is
∘
Δ𝐻rxn,𝑇𝑟𝑒𝑓
= ∑ 𝜈𝑖 Δ𝐻𝑓∘𝑖,𝑇𝑟𝑒𝑓
𝑖
For species in the most stable state (e.g. monatomic carbon or diatomic oxygen), the enthalpies of formation
are assigned to be zero.
The Δ𝐶𝑃∘ term is change heat capacity term weighted by stoichiometric ratios and is also a function of
temperature. This is
∘
Δ𝐶𝑃∘ (𝑇) = ∑ 𝜈𝑖 𝐶𝑃,𝑖
𝑖
∘
The value of the heat capacity for a given species (𝐶𝑃,𝑖 ) can be found in reference tables as well. It is often
reported as a power series of temperature with empirically derived coefficients.
An analogous expression can be written for the standard Gibbs free energy change of reaction at a reference
temperature:
∘
Δ𝐺rxn,𝑇𝑟𝑒𝑓
= ∑ 𝜈𝑖 Δ𝐺𝑓∘𝑖,𝑇𝑟𝑒𝑓
𝑖
∘ ∘
However, it is critical to note that since Δ𝐺rxn are typically found from Δ𝐺𝑓,𝑖 at a reference temperature,
𝑇ref, then the resulting 𝐾𝑎 is only valid at 𝑇ref and is
∘
𝛥𝐺rxn,𝑇𝑟𝑒𝑓
𝐾𝑎,𝑇𝑟𝑒𝑓 = exp (− )
𝑅𝑇𝑟𝑒𝑓
In order to determine the 𝐾𝑎 at a temperature other than the reference temperature, one must use the
following relationship
𝑇2
𝐾𝑎,𝑇2 ∘ (𝑇)
Δ𝐻rxn
ln ( )= ∫ 𝑑𝑇
𝐾𝑎,𝑇1 𝑅𝑇 2
𝑇1
∘
where 𝑇1 typically is 𝑇ref. Since Δ𝐻rxn is a function of temperature, this must usually be accounted for
using the aforementioned temperature-dependence relationships. However, if this temperature effect can be
ignored, then the equation simplifies to
𝐾𝑎,𝑇2 Δ𝐻∘rxn 1 1
ln ( )=− ( − )
𝐾𝑎,𝑇1 𝑅 𝑇2 𝑇1
𝑃 𝛿 𝜈
𝐾𝑎 = ( ) ∏ 𝑦𝑖 𝑖
𝑃ref
𝑖
1
Since partial pressure is just 𝑃𝑖 = 𝑃𝑦𝑖 , we can optionally rewrite activity as 𝑎𝑖 = 𝑃𝑖 /𝑃ref
CONSEQUENCES OF CHEMICAL EQUILIBRIA | 20
where 𝛾𝑖 is specifically the Raoult’s law activity coefficient and indicates ideal behavior according to
Raoult’s law. As 𝑥𝑖 → 1, the value of the activity coefficient also approaches 1. As 𝑥𝑖 → 0, the activity
coefficient approaches an infinitely dilute case denoted 𝛾𝑖∞ and can be found by divided the Henry’s law
constant by the saturation vapor pressure. The value of 𝛾𝑖∞ is large for poorly soluble, low volatility species.
∞ ∞
If 𝛾product > 𝛾reactants then there is an equilibrium shift toward products.
Dilute solutions typically follow Henry’s law instead of Raoult’s law, and it is more conventional to use
concentration. As such,
𝐶𝑖
𝑎𝑖 = 𝛾𝑖
𝐶ref
where 𝐶𝑖 refers to the concentration of a species 𝑖, 𝐶ref is typically 1 M. For an ideal liquid mixture, the
activity coefficient approaches 1 and
𝐶𝑖
𝑎𝑖 =
𝐶ref
Therefore, for ideal solutions
1 𝛿 𝜈
𝐾𝑎 = ( ) ∏ 𝐶𝑖 𝑖
𝐶ref
𝑖
𝜈
𝐾𝐶 = ∏ 𝐶𝑖 𝑖
𝑖
𝜈𝑖
𝐾𝑦 = ∏ 𝑦𝑖
𝑖
1 𝛿 𝑃 𝛿
𝐾𝑎 = ( ) 𝐾𝑃 = ( ) 𝐾𝑦
𝑃ref 𝑃ref
𝐾𝑃 = 𝐾𝐶 (𝑅𝑇)𝛿
For an ideal liquid mixture, we have that
1 𝛿
𝐾𝑎 = ( ) 𝐾𝐶
𝐶ref
We therefore see that for ideal solutions, 𝐾𝑎 ≈ 𝐾𝑃 ≈ 𝐾𝐶 (taking into account reference terms to take care
of units).
CONSEQUENCES OF CHEMICAL EQUILIBRIA | 21
The value of 𝐾M is the concentration of A at which the rate is half of its maximum. The value of 𝑉max is the
maximum rate at [A] ≫ 𝐾M . From this equation, it is clear that at [A] ≪ 𝐾M , the rate appears to be first
order in A. At [A] ≫ K 𝑀 , the rate is pseudo zeroth order in A (and therefore independent of A). At all [A],
the rate is proportional to the total enzyme concentration.
CONSEQUENCES OF CHEMICAL EQUILIBRIA | 22
The Eadie-Hofstee plot is another way to linearize the Michaelis-Menten equation, which uses the following
equation
𝑟P
𝑟P = −𝐾M + 𝑉max
[𝐴]
and has a plot with characteristics shown below
Since the Lineweaver-Burke plot uses inverse values, it unevenly weights data points in concentration and
reaction rate whereas Eadie-Hofstee plots do not. However, the Eadie-Hofstee plot’s abscissa and ordinate
are dependent on the reaction rate, and therefore any experimental error will be present in both axes. Any
uncertainty also propagates unevenly and becomes larger over the abscissa, giving weight to smaller values
of 𝑟P /[A].
In a batch reactor, one often wants to plot [A] as a function of time. To obtain this, the Michaelis-Menten
equation must be integrated. Doing so yields
𝐶𝐴0
𝐾M ln ( ) + (𝐶𝐴0 − 𝐶𝐴 ) = 𝑘3 𝐸T 𝑡
𝐶𝐴
CONSEQUENCES OF CHEMICAL EQUILIBRIA | 23
Plotting concentration as a function of time yields a function that clearly changes slope as it switches from
first order to zeroth order regimes
𝐶𝐴0 − 𝐶𝐴 𝑡
= 𝑉max ( ) − 𝐾M
𝐶 𝐶
ln ( 𝐶𝐴0 ) ln ( 𝐶𝐴0 )
𝐴 𝐴
In the case of no reversible binding, at 𝑡 → ∞ we will have [A] → 0. However, for reversible binding, at
𝑡 → ∞ we will have [A] approach some constant non-zero equilibrium value.
CONSEQUENCES OF CHEMICAL EQUILIBRIA | 24
6 REACTION NETWORKS
6.1 INTRODUCTION TO REACTION NETWORKS
Previously, we have discussed rate laws for a single reaction of the type 𝐴 → 𝐵. However, oftentimes we
have to deal with a complex reaction network, where there are many products produced from a reactant (or
multiple reactants), but little information is known a priori about how the products are actually produced.
Specifically, the reaction topology is often desired. In this section, we will refer to the “rank” (denote 𝑟) of
a species as the numerical order in which a product appears. If it comes first, it will have rank 1 and be
called primary; if second, it will have rank 2 and be called secondary, and so on. This is distinct from the
order of that species, which we will reserve for the value of 𝑛 in rate = 𝑘𝐶𝐴𝑛 (i.e. the dependence of rate
on concentration).
We will define yield of a species 𝑖 (compared to a reactant 𝐴) as
𝑁𝑖 − 𝑁𝑖0
𝑌𝑖 ≡
𝑁𝐴0
We will define the conversion of a reactant 𝐴 as
𝑁𝐴
𝑋𝐴 ≡ 1 −
𝑁𝐴0
We will define selectivity of a species 𝑖 (compared to a reactant 𝐴) as
𝑌𝑖 𝑁𝑖 − 𝑁𝑖0
𝑆𝑖 ≡ =
𝑋𝐴 𝑁𝐴0 − 𝑁𝐴
The differential selectivity will therefore be defined as
𝜕𝑌𝑖
𝑆𝑑𝑖 ≡
𝜕𝑋𝐴
6.2 DELPLOTS
To determine the rank of a reaction product, the method of delplots can be used. The first-rank delplot is
used to distinguish primary products from non-primary products. Higher rank delplots allow for the
discernment of products that are secondary, tertiary, and so on. The following notation will be used to
describe delplots: 1PA . Here, the left-hand superscript is a number representing the rank of the delplot (in
this case 1), P represents a product P, and the subscript A means that it is based on the conversion of species
A.
The first-rank delplot is a plot of selectivity versus conversion. In other words, it is a plot of 𝑌𝑃 /𝑋𝐴 vs. 𝑋𝐴 .
For each product P, the delplot is extrapolated to 𝑋𝐴 → 0 (i.e. when 𝑡 → 0 or 𝜏 → 0). This y-intercept,
denoted as 1PA , can be evaluated as
1 𝑃/𝐴0 𝑃
PA = lim = lim
𝜏→0 (𝐴0 − 𝐴)/𝐴0 𝜏→0 𝐴0 − 𝐴
This can be evaluated algebraically from given expressions of the concentrations of each product as a
function of time or numerically from the actual plots of selectivity versus conversion.
REACTION NETWORKS | 26
If a finite y-intercept is found in the first-rank delplot, then the product is primary because the initial rate
of formation of a primary product is always finite. If the y-intercept is zero, then the product is non-primary
(i.e. the rank is greater than 1) because the initial rate of a non-primary product is always zero (i.e. it lags).
Higher rank delplots allows for the sorting of products of rank greater than 1. The second-rank delplot
consists of a plot of 𝑌𝑃 /𝑋𝐴2 vs. 𝑋𝐴 . The y-intercept of the second-rank delplot for a product P is
2 𝑃/𝐴0 𝐴0 𝑃
PA = lim 2 = lim
𝜏→0 𝐴 −𝐴 𝜏→0 (𝐴0 − 𝐴)2
( 0𝐴 )
0
If a finite intercept is found, the product is secondary. If a zero intercept is found, the product’s rank is
greater than secondary. If a divergence is found as 𝜏 → 0 (i.e. no y-intercept), the product’s rank is primary.
This can be summarized generally. The delplot of rank 𝑚 is a plot of 𝑌𝑃 /(𝑋𝐴 )𝑚 vs. 𝑋𝐴 . The y-intercept can
be found via
𝑚 𝑃/𝐴0
PA = lim
𝜏→0 𝐴0 − 𝐴 𝑟
( )
𝐴0
The following general procedure can then be employed, which has been extended to include orders other
than first order. In the table, 𝑟 is the rank of the species and 𝑚 is the rank of the plot. Note that it is often
too difficult to notice changes from 0 intercepts to finite intercepts for 𝑛 > 1 when the rank is high. If you
encounter a species that has a 0 intercept no matter the rank of the Delplot, it is likely either 𝑛 = 1 and a
really high rank species (perhaps it is produced from intermediate species you are not aware of) or 𝑛 > 1
and some rank at the end of the chain. The specific order can often be determined by looking at
concentration versus time plots. In general, Delplots are limited at ranks greater than about 2 or 3 due to
the large propagations of error that occur.
7 KINETIC THEORY
7.1 COLLISION THEORY
7.1.1 DISTRIBUTION LAWS
Collision theory can be used to estimate kinetic parameters from first-principles. It assumes that the
molecules are hard spheres and react by collision with one another. It also assumes that the attractive forces
between the species are negligible.
Although it will not be derived in this guide, it can be found that the probability distribution for a given
speed, 𝑐, is given by
4𝑐 2 𝑐2
𝑃(𝑐) 𝑑𝑐 = exp (− ) 𝑑𝑐
𝛼 3 √𝜋 𝛼2
2𝑘𝐵 𝑇
𝛼=√
𝑚
The distribution of kinetic energy can be derived from this expression and can be given by
4 2𝐸 2𝐸
𝑃(𝐸) 𝑑𝐸 = 3
√ 3 exp (− ) 𝑑𝐸
𝛼 𝜋𝑚 𝑚𝛼 2
8𝑘𝐵 𝑇
𝑐avg = √
𝜋𝜇𝐴𝐵
For the case of identical molecules, the reduced mass simplifies to half the mass of the molecule. This
relative speed is useful because now we can write the collision frequency of 𝐴 moving through a matrix of
𝐵 molecules when 𝐴 has a Maxwell distribution of speeds
2
8𝑘𝐵 𝑇
𝑍𝐴𝐵 = 𝜋𝜎𝐴𝐵 𝑛𝐵 √
𝜋𝜇𝐴𝐵
2
8𝑘𝐵 𝑇
𝑍𝐴𝐵 = 𝜋𝜎𝐴𝐵 𝑛𝐴 𝑛𝐵 √
𝜋𝜇𝐴𝐵
If there is only one molecular species 𝐴 then the collision frequency is given by
2 2√ 𝐵
𝑘 𝑇
𝑍𝐴𝐴 = 4𝜋𝜎𝐴𝐴 𝑛𝐴
𝜋𝑚𝐴
the pure collision of two species, it not reflective of the kinetics of reactions because there is no exponential
dependence on temperature. The missing link here is that molecules only react if they collide with sufficient
energy transfer. We therefore must define a threshold energy, 𝜂 ∗ , above which reactions can occur. We then
find that the reactive collision number is
2 √
8𝑘𝐵 𝑇 𝜂∗
𝑍 = 𝑝𝜋𝜎𝐴𝐵 exp (− )𝑛 𝑛
𝜋𝜇𝐴𝐵 𝑘𝐵 𝑇 𝐴 𝐵
Here, 𝑝 is a steric factor which is effectively a correction factor (unless otherwise defined, it can be taken
as a value of 1). For 𝑟𝐴 = −𝑘𝑛𝐴 𝑛𝐵 , we then have that
2 √
8𝑘𝐵 𝑇 𝜂∗
𝑘 = 𝑝𝜋𝜎𝐴𝐵 exp (− )
𝜋𝜇𝐴𝐵 𝑘𝐵 𝑇
From the above expression, we see that there is an exponential temperature dependence as would be
expected from Arrhenius’ law, but there is also a √𝑇 dependence in what would be analogous to the pre-
exponential factor, which is otherwise assumed to be independent of temperature in the Arrhenius equation.
The collision theory equation is fairly accurate for low values of 𝜂 ∗ , such as very exothermic elementary
steps like radical recombination.
𝑘1 ,𝑘−1
𝐴+𝐴↔ 𝐴 + 𝐴∗
𝑘2 ,𝑘−2
𝐴∗ ↔ products
If we apply PSSH on the activated intermediate, we get
𝑑𝐶𝐴∗
= 0 = 𝑘1 𝐶𝐴2 − 𝑘−1 𝐶𝐴 𝐶𝐴∗ − 𝑘2 𝐶𝐴∗
𝑑𝑡
Solving for 𝐶𝐴∗ yields
𝑘1 𝐶𝐴2
𝐶𝐴∗ =
𝑘2 + 𝑘−1 𝐶𝐴
The rate of product formation is then given by
𝑘1 𝑘2 𝐶𝐴2
𝑟 = 𝑘2 𝐶𝐴∗ =
𝑘2 + 𝑘−1 𝐶𝐴
At high pressures (i.e. high concentrations), we get
𝑘1 𝑘2
𝑟≈ 𝐶 = 𝑘∞ 𝐶𝐴
𝑘−1 𝐴
where 𝑘∞ is the observable rate constant. At low pressures (i.e. low concentrations) we get
𝑟 ≈ 𝑘1 𝐶𝐴2
We know that collision theory works very well for radical decomposition, so we can assume that 𝑘−1 can
be estimated well from collision theory and is therefore
2 2√
8𝑘𝐵 𝑇
𝑘−1 = 𝜋𝜎𝐴𝐴 𝑛𝐴
𝜋𝜇𝐴𝐵
where the exponential energy term can be ignored because 𝜂 ∗ ≈ 0 for this reaction. The probability
distribution function of energies in this system can also be found to be
𝐸∗ 𝐸 ∗ 𝑛−1
exp (− )( )
𝑘𝐵 𝑇 𝑘𝐵 𝑇
𝑓(𝐸 ∗ ) =
(𝑛 − 1)!
where 𝑛 is the number of degrees of vibrational freedom. For a nonlinear polyatomic molecule, it is given
by 𝑛 = 3𝑁 − 6 and a linear molecule it is given by 𝑛 = 3𝑁 − 5 where 𝑁 is the number of atoms in the
molecule. We can say that the equilibrium constant for the first reaction is given by
𝑘1 𝐶𝐴∗
𝐾1 = =
𝑘−1 𝐶𝐴
The term 𝐶𝐴∗ /𝐶𝐴 is essentially the fraction of the total number of molecules with energy greater than 𝐸 ∗ , so
𝑘1
= 𝑓(𝐸 ∗ )
𝑘−1
KINETIC THEORY | 30
𝐸∗ 𝐸 ∗ 𝑛−1
8𝑘𝐵 𝑇 exp (− ) ( )
2 2√ 𝑘𝐵 𝑇 𝑘𝐵 𝑇
𝑘1 = 𝜋𝜎𝐴𝐴 𝑛𝐴
𝜋𝜇𝐴𝐵 (𝑛 − 1)!
If we look back at the rate constant for the high concentration case,
𝑘1 𝑘2
𝑘∞ = = 𝑘2 𝑓(𝐸)
𝑘−1
If we want to know what the observable rate constant is, we need to find 𝑘2 . The value of 𝑘2 is the
probability that the product decays and is given by
𝐸 ∗ 1−𝑛
𝑘2 = 𝜈̅ (1 − )
𝑛𝑘𝐵 𝑇
After a lot of math that is omitted here, we can arrive at the following expression for the observable rate
constant in the limit of high 𝐸 ∗ , which is
𝐸∗
𝑘∞ = 𝜈̅ exp (− )
𝑘𝐵 𝑇
which has the familiar form of the Arrhenius expression (𝜈̅ is an intrinsic frequency factor and thus has the
units of the pre-exponential factor in the Arrhenius equation). We then see that Lindemann theory succeeds
where collision theory fails – in the high 𝐸 ∗ limit.
[𝐴𝐵𝐶 ‡ ]
𝐾𝐶‡ =
[𝐴𝐵][𝐶]
This can be determined from statistical mechanics via the following expression2
𝑁𝐴𝑛−1 𝑄 ‡ 𝛥𝑈 ‡
𝐾𝐶‡ = exp (− )
𝑄𝐴𝐵 𝑄𝐶 𝑅𝑇
where 𝑛 is the molecularity. The Δ𝑈 ‡ can only be obtained from calculations (typically the difference of
zero point energies of reactants and the transition state), the value of 𝑛 is the molecularity of the reaction,
𝛥𝑈 ‡ 𝜈
2
More generally, for the reaction 𝑎𝐴 + 𝑏𝐵 ↔ 𝑐𝐶 + 𝑑𝐷, we have that 𝐾𝐶 = 𝑁𝐴𝑛−1 exp (− 𝑅𝑇
) ∏𝑖 𝑄𝑖 𝑖 .
KINETIC THEORY | 31
and 𝑁𝐴 is Avogadro’s number (to convert the units properly to a mole basis). The 𝑄 values here represent
partition functions with units of inverse volume and are the products of vibrational, rotational, translational,
and electronic partition functions. Written mathematically,
𝑄 = 𝑄𝑡 𝑄𝑟 𝑄𝑣𝑁 𝑄𝑒
where 𝑁 is the number of vibrational modes. It is 3𝑁0 − 5 for linear molecules and 3𝑁0 − 6 for nonlinear
polyatomic molecules where 𝑁0 is the number of atoms. However, a key point must be addressed for
transition states. The power of 𝑁 must be decreased by one for transition states because there is one less
degree of freedom due to the pseudo-bond formed. For this,
𝑄 ‡ = 𝑄𝑡 𝑄𝑟 𝑄𝑣𝑁−1 𝑄𝑒
We now must define each of the partition functions. The translational partition function is defined as
3
2𝜋𝑚𝑘𝐵 𝑇 2
𝑄𝑡 ≡ ( )
ℎ2
and has units of inverse volume. It typically has a value of about 1024 cm-3. Importantly, if we can constrain
the system to be in a 2D space instead of 3D space, the exponent drops by a factor of 1/2.
The rotational partition function is different depending on the shape of the molecule. For a linear molecule,
8𝜋 2 𝐼𝑘𝐵 𝑇
𝑄𝑟 =
𝜎ℎ2
and for a nonlinear molecule
1 3
8𝜋 2 (8𝜋 3 𝐼1 𝐼2 𝐼3 )2 (𝑘𝐵 𝑇)2
𝑄𝑟 =
ℎ3 𝜎
In these equations, 𝜎 represents the symmetry number and is determined by the number of spatial
orientations of the subject molecule that are identical. For easy reference, it is a value of 2 for linear
molecules with a center of symmetry and 1 for linear molecules without a center of symmetry. The quantity
𝐼 is the moment of inertia, and for the nonlinear case they are the three principal moments. The moment of
inertia is defined as
𝐼 = ∑ 𝑚𝑖 𝑟𝑖2
𝑖
where 𝑟 is the distance to the axis of rotation. For a diatomic molecule, the moment of inertia is 𝐼 = 𝜇𝑅 2
where 𝜇 is the reduced mass and 𝑅 is the distance between the two atoms. For a linear, symmetric molecule
2
like CO2, the moment of inertia is 𝐼 = 2𝑚O 𝑟CO , where 𝑚O is the mass of the oxygen atom and 𝑟CO is the
C-O bond length. For a triatomic linear molecule, such as a D-D-H transition state, the moment of inertia
2 2
would approximately be 𝐼 = 𝑚𝐷 𝑟𝐷𝐷 + 𝑚𝐻 𝑟𝐻𝐷 (this assumes the center of the molecule is the center of
mass, which is a reasonable approximation). The value of 𝑄𝑟 is unitless and approximately 102 − 104 for
linear molecules and 103 − 106 for nonlinear molecules.
The vibrational partition function is given by
KINETIC THEORY | 32
𝑛
ℎ𝜈𝑖 −1
𝑄𝑣 = ∏ (1 − exp (− ))
𝑘𝐵 𝑇
𝑖
where 𝑛 is the degrees of vibrational freedom and 𝜈𝑖 is the vibrational frequency from IR or Raman
spectroscopy. The value of 𝑄𝑣 is unitless and approximately 1 to 10. Note that spectra normally yield
wavenumbers with units of inverse length. To convert a wavenumber 𝜈̃ to frequency, use 𝜈𝑖 = 𝑐𝜈̃𝑖 .
Finally, the electronic partition function is given by
𝐸𝑖
𝑄𝑒 = ∑ 𝑔𝑖 exp (− )
𝑘𝐵 𝑇
𝑖
where 𝑔𝑖 is the degeneracy and 𝐸𝑖 is the electronic energy above the ground state.
If it is impossible or too inconvenient to calculate the partition functions directly, one can still make order
of magnitude arguments. Generally, we can say that
𝑄𝑖 ≈ 𝑓𝑡3 𝑓𝑟𝑎 𝑓𝑣𝑁 𝑓𝑒
where 𝑎 = 2 and 𝑁 = 3𝑁0 − 5 for a linear molecule and 𝑎 = 3 and 𝑁 = 3𝑁0 − 6 for a polyatomic
nonlinear molecule. For the transition state,
We then can plug in 𝑓𝑡3 ≈ 1024 cm−3, 𝑓𝑟2 ≈ 102 − 104 (linear) or 𝑓𝑟3 ≈ 103 − 106 (nonlinear), 𝑓𝑣 ≈ 1 −
3, and 𝑓𝑒 ≈ 1.
When using TST, oftentimes a number of approximations are made to make the math easier. Generally
speaking, 𝑄𝑒 is neglected unless dealing with transition metals in their excited states. Also, for species with
large vibrational frequencies, 𝑄𝑣 ≈ 1. Another common thing to realize is that 𝑄𝑟 = 1 if the species is
monatomic. When dealing with a heterogeneous catalyst surface, the following assumptions can typically
be made: the surface has all partition functions being approximately 1, and the transition state has 𝑄𝑡 ≈ 1.
𝑟 = 𝑣‡ [𝐴𝐵𝐶 ‡ ]
where 𝑣‡ is the frequency of passage over the energy barrier and [𝐴𝐵𝐶 ‡ ] is the concentration of the
transition state. Substituting in for 𝐶‡ using the previously defined 𝐾𝐶‡ yields
𝑟 = 𝑣‡ 𝐾𝐶‡ [𝐴𝐵][𝐶]
We can define the frequency 𝑣‡ as the thermal energy provided such that
𝑅𝑇 𝑘𝐵 𝑇
𝑣‡ = =
𝑁𝐴 ℎ ℎ
so that we can say
𝑘𝐵 𝑇 ‡
𝑟= 𝐾 [𝐴𝐵][𝐶]
ℎ 𝐶
KINETIC THEORY | 33
𝑘𝐵 𝑇 𝑁𝐴𝑛−1 𝑄 ‡ 𝛥𝑈 ‡
𝑟= exp (− ) [𝐴𝐵][𝐶]
ℎ 𝑄𝐴𝐵 𝑄𝐶 𝑅𝑇
We should note that this takes the form of an Arrhenius-like expression. We can say that the pre-exponential
factor is
𝑘𝐵 𝑇 𝑁𝐴𝑛−1 𝑄 ‡
𝑘0 =
ℎ 𝑄𝐴𝐵 𝑄𝐶
such that
𝛥𝑈 ‡
𝑘 = 𝑘0 exp (− )
𝑅𝑇
Δ𝐺 ∘‡ = −𝑅𝑇 ln(𝐾𝑎‡ )
Δ𝐺 ∘‡ = Δ𝐻 ∘‡ − 𝑇Δ𝑆 ∘ ‡
to say
𝛥𝑆 ∘ ‡ 𝛥𝐻 ∘‡
𝐾𝑎‡ = exp ( ) exp (− )
𝑅 𝑅𝑇
𝑘𝐵 𝑇 𝛥𝑆 ∘ ‡ 𝛥𝐻 ∘‡
exp ( 𝑅 ) exp (− 𝑅𝑇 )
ℎ
𝑟= [𝐴𝐵][𝐶]
𝐶 ∘ 𝑛−1
where 𝐶 ∘ is a reference concentration (usually given by 𝑃/𝑅𝑇 at 1 atm and 𝑇 for gases or as the molal
concentration of pure components for liquids) and 𝑛 is the molecularity.
KINETIC THEORY | 34
The Hammett equation relates the substituted group 𝑥 with the rate of reaction and takes into account the
effect of inductive and resonance effects. It can be given by
𝑘𝑥
log ( ) = 𝛾𝐻 𝜎𝑥
𝑘𝐻
𝐾
where 𝜎𝑥 ≡ log (𝐾𝑥 ) and is a tabulated quantity3. The lower case 𝑘 values refer to reaction rate constants,
𝐻
and the capital 𝐾 values refer to equilibrium (acid) dissociation constants. For the Hammett equation, a
substituted group of 𝑥 = 𝐻 (i.e. the group is a hydrogen atom) is often used as a reference.
Higher values of 𝜎𝑥 indicate that the equilibrium constant is increased for a chemical group 𝑥 with respect
to the hydrogen reference, meaning that the reaction is shifted more in the forward direction (specifically,
via a stabilization of negative charge via resonance and inductive effects). Therefore, the acidity of the
reactant molecule increases when 𝜎𝑥 > 0. Such a substituent is considered to be an electron-withdrawing
group because it will stabilize the product anion by withdrawing negative charge away from the reaction
site. Conversely, molecules with 𝜎𝑥 < 0 indicate an electron-donating group where the acid dissociation is
disfavored since the electron density is increased near the reaction site.
There are also trends that can be seen with the 𝛾𝐻 values, which are easier to see when you make the
equation 𝑘𝑥 = 𝑘𝐻 10𝛾𝐻 𝜎𝑥 . For 𝛾𝐻 > 1, the reaction is more sensitive to substituents than the reference
molecule, and negative charge is built up in the transition state of the reaction. The reaction is assisted by
electron-withdrawing groups. For 𝛾𝐻 = 0, there is no sensitivity to substituents, and no charge is built or
lost. For 𝛾𝐻 < 0, the reaction builds up positive charge in the transition state of the reaction. The reaction
is assisted by electron-donating groups.
It should also be noted that 𝜎𝑥 values are approximately additive, so if an aromatic ring has 2 Cl groups,
then the 𝜎𝑥 value can be approximated as 2𝜎Cl .
There are alternate forms of the Hammett parameter that are parameterized for the reactions that require
positive and negative charge stabilization, given by 𝜎 + and 𝜎 − , respectively. For pure inductive effects, 𝜎′
can be used.
3
Some references use log10 whereas others use ln.
KINETIC THEORY | 35
where 𝜎 ∗ is a purely inductive effect and 𝐸𝑠 is a purely steric effect. In the limit of steric effects being
negligible, it simplifies to the same functional form as the Hammett equation. The trends in 𝛾 ∗ are then the
same as the trends for 𝛾𝐻 .
SURFACE CATALYSIS | 36
8 SURFACE CATALYSIS
8.1 ADSORPTION RATE LAWS
8.1.1 MOLECULAR ADSORPTION
We will now consider surface catalysis, wherein a reactant in a continuum diffuses into a catalyst particle,
binds to the surface, reacts to form a new species, desorbs, diffuses out of the catalyst particle, and enters
the continuum. As can be seen from this outline, heterogeneous catalysis incorporates adsorption, transport
phenomena, reactions, and bulk reactor models. We will focus on developing a basic kinetic model first
from elementary steps including adsorption, desorption, and surface reaction.
Consider the first step: adsorption. We will use an asterisk to denote a surface site. As such, the adsorption
step can be given by
𝑘1 ,𝑘−1
𝐴 +∗↔ 𝐴∗
For example, this could be CO + S ∗ ↔ CO∗ (note that when dealing with gases, such as CO, it is standard
to use a partial pressure instead of concentration). This specific type of adsorption is referred to molecular
(or non-dissociative) adsorption. The equation for the net rate of adsorption can then be written as
𝐶𝐴∗
𝑟ad = 𝑘1 𝐶𝐴 𝐶∗ − 𝑘−1 𝐶𝐴∗ = 𝑘1 (𝐶𝐴 𝐶∗ − )
𝐾𝐴
where 𝐾𝐴 ≡ 𝑘1 /𝑘−1 and is referred to as the adsorption equilibrium constant. The adsorption rate constant
𝑘1 is relatively independent of temperature whereas the desorption constant 𝑘−1 increases exponentially
with increasing temperature, such that the equilibrium adsorption constant 𝐾𝐴 decreases exponentially with
increasing temperature. A site balance can generally be written as
𝐶𝑇 = 𝐶∗ + ∑ 𝐶𝑖∗
𝑖
where the summation term accounts for any other adsorbed species to the surface sites. In this example,
𝐶𝑇 = 𝐶∗ + 𝐶𝐴∗
We know that at equilibrium, the net rate of adsorption should be zero. Employing this condition yields
𝐶𝐴∗ = 𝐾𝐴 𝐶𝐴 𝐶∗
Substituting in for 𝐶∗ from the site balance yields
𝐶𝐴∗ = 𝐾𝐴 𝐶𝐴 (𝐶𝑇 − 𝐶𝐴∗ )
which can be rearranged to
𝐾𝐴 𝐶𝐴 𝐶𝑇
𝐶𝐴∗ =
1 + 𝐾𝐴 𝐶𝐴
This equation is specifically called the Langmuir isotherm. This equation can be linearized to the following
form for data-fitting purposes, as shown below
𝐶𝐴 1 𝐶𝐴
= +
𝐶𝐴∗ 𝐾𝐴 𝐶𝑇 𝐶𝑇
SURFACE CATALYSIS | 37
Typically, the parameter 𝜃 is defined as the number of moles of species adsorbed divided by the number of
moles in the monolayer. This would be
𝐶𝐴∗ 𝐾𝐴 𝐶𝐴
𝜃= =
𝐶𝑇 1 + 𝐾𝐴 𝐶𝐴
The linearized form is
1 1
= +1
𝜃 𝐾𝐴 𝐶𝐴
Oftentimes, 𝜃 is defined as the volume adsorbed divided by the volume adsorbed at saturation: 𝑉/𝑉𝑀 . In
this definition, the above expression simply becomes
1 1 1 1
= ( )( )+
𝑉 𝐶𝐴 𝐾𝐴 𝑉𝑀 𝑉𝑀
This could be CO + 2 ∗↔ C ∗ + O∗ , for instance. In this case, the adsorption rate can be written as
𝐶𝐴∗ 𝐶𝐵∗
𝑟ad = 𝑘1 𝐶𝐴𝐵 𝐶∗2 − 𝑘−1 𝐶𝐴∗ 𝐶𝐵∗ = 𝑘1 (𝐶𝐴𝐵 𝐶∗2 − )
𝐾𝐴
where 𝐾𝐴 ≡ 𝑘1 /𝑘−1 once again. For dissociative adsorption, both 𝑘1 and 𝑘−1 increase exponentially with
temperature, unlike the case with molecular adsorption. We can then employ equilibrium conditions such
that
1
(𝐾𝐴 𝐶𝐴𝐵 )2
𝜃= 1
1 + 2(𝐾𝐴 𝐶𝐴𝐵 )2
The equation can be linearized by
√𝐶𝐴𝐵 1 2√𝐶𝐴𝐵
= +
𝐶𝐴∗ 𝐶𝑇 √𝐾𝐴 𝐶𝑇
which equivalently is the following when 𝐶𝑇 is multiplied through and √𝐶𝐴𝐵 is divided through
1 1
= +2
𝜃 √𝐾𝐴 𝐶𝐴𝐵
It can be easily shown that the dissociative adsorption process of A2 + 2 ∗↔ 2𝐴∗ is identical to the case
1
before but the denominator in the expression for 𝐶𝐴∗ is 1 + (𝐾𝐴 𝐶𝐴2 )2 where the factor of 2 disappears
because the site balance is just 𝐶𝑇 = 𝐶∗ + 𝐶𝐴∗ . Therefore, in this case
1
(𝐾𝐴 𝐶𝐴2 )2 𝐶𝑇
𝐶𝐴∗ = 1
1 + (𝐾𝐴 𝐶𝐴2 )2
Oftentimes, a plot of 1/𝜃 vs. 1/√𝐶𝐴2 will have deviations at high and low coverages (and by extension
high and low partial pressures) due to adsorbate-adsorbate interactions and site heterogeneity.
We know that
𝐶𝐵∗ = 𝐾𝐵 𝐶𝐵 𝐶∗
So
𝐶𝐴∗ = 𝐾𝐴 𝐶𝐴 (𝐶𝑇 − 𝐶𝐴∗ − 𝐾𝐵 𝐶𝐵 𝐶𝐴∗ )
which can be rearranged to
𝐾𝐴 𝐶𝐴 𝐶𝑇
𝐶𝐴∗ =
1 + 𝐾𝐴 𝐶𝐴 + 𝐾𝐵 𝐶𝐵
An analogous procedure for B would yield
𝐾𝐵 𝐶𝐵 𝐶𝑇
𝐶𝐵∗ =
1 + 𝐾𝐴 𝐶𝐴 + 𝐾𝐵 𝐶𝐵
8.5 NONIDEALITIES
8.5.1 NONIDEAL SURFACES
Real surfaces are not perfect and have defects/deformations that cause Δ𝐻ads to not be constant over the
entire surface. We know that the equilibrium constant is
𝛥𝐺𝑎𝑑𝑠 𝛥𝑆𝑎𝑑𝑠 Δ𝐻ads
𝐾eq = exp (− ) = exp ( − )
𝑅𝑇 𝑅 𝑅𝑇
From thermodynamics and the above expression, we then see that Δ𝐻ads < 0 is favorable and Δ𝑆ads > 0
is favorable. In reality, Δ𝐻ads is a function of 𝜃. The Langmuir models, derived in the following subsections,
assume Δ𝐻ads is a constant and is only good for isolated sites.
Additional models have been developed. The Temkin model states that
0 (1
Δ𝐻ads (𝜃) = Δ𝐻𝑎𝑑𝑠 − 𝛼𝜃)
0
where Δ𝐻ads is the heat of adsorption at zero surface coverage and 𝛼 is a fitting parameter. This expression
is an empirical relationship that simply makes Δ𝐻ads linearly decrease with increasing 𝜃. For molecular
adsorption, the Temkin isotherm is
𝑅𝑇
𝜃≈ 0 ln(𝐾𝐴 𝐶𝐴 )
Δ𝐻ads 𝛼
This model completely breaks down as 𝜃 → 0. It can be shown via the Clausius-Clapeyron equation that
the Freundlich isotherm for molecular adsorption takes the form of
𝜃 = 𝛼𝐶𝐴𝑚
where 𝛼 and 𝑚 fitting parameters. The value of 𝑚 is directly related to the intensity of adsorption. The
value of 𝛼 is an indicator of the adsorption capacity.
This is showing that a diatomic gas A2 comes in contact with the surface but does not always bind it. We
can apply PSSH to the A2,surface intermediate such that
𝑑[𝐴2,surfrace ]
= 𝜙𝐹𝐴 − 𝑘𝑑 [𝐴2,surface ] − 𝑘𝑎 [𝐴2,surface ](1 − 𝜃)2 ≈ 0
𝑑𝑡
SURFACE CATALYSIS | 42
Here, 𝜙 is the trapping ratio, 𝐹𝐴 is the flux of 𝐴, and 𝜃 is the fractional coverage. Note that the (1 − 𝜃)2
term comes in because that is effectively 𝐶∗2 /𝐶𝑇2 . We know that the sticking coefficient is
At zero coverage,
𝑘𝑎 𝜙
𝑆(0) =
𝑘𝑑 + 𝑘𝑎
so that
𝑆(𝜃) (𝑘𝑑 + 𝑘𝑎 )(1 − 𝜃)2
=
𝑆(0) 𝑘𝑑 + 𝑘𝑎 (1 − 𝜃)2
If we define 𝐾 ≡ 𝑘𝑎 /𝑘𝑑 then we arrive at
𝑆(𝜃) (1 + 𝐾)(1 − 𝜃)2
=
𝑆(0) 1 + 𝐾(1 − 𝜃)2
Recall from the molecular adsorption section that 𝐾 decreases exponentially with increasing temperature.
As such, at the low temperature limit and low values of 𝜃, we find that 𝑆(𝜃)/𝑆(0) ≈ 1. At the high
temperature limit, we find that 𝑆(𝜃)/𝑆(0) ≈ (1 − 𝜃)2.
REACTIONS IN HETEROGENEOUS SYSTEMS | 43
where 𝑊 is an arbitrary volume/surface area expression for the particular shape. It should be noted that this
definition of 𝜙𝑛2 is different by a numerical factor. Generally, just use the first definition for spheres and
the second one for more complicated geometries. Everything is an order of magnitude argument anyway.
4
Note that the definition of the Thiele modulus oftentimes has the 𝑘𝑛 term replaced with a 𝑘𝑛 𝜌𝐵 𝑆𝐴 term. In this, 𝜌𝐵
is the bulk density and 𝑆𝐴 is the surface area divided by the mass of the catalyst. This is because oftentimes rates are
reported in moles/(cm2 s) and therefore 𝑘𝑛 has different units than usual. If one is dealing with a rate that has
moles/(cm3 s) or equivalent dimensions, then simply omit the 𝑆𝐴 𝜌𝐵 unit correction term in the definition. If one is
dealing with a rate that has units of 1/s then keep the 𝜌𝐵 term but omit the 𝑆𝐴 term.
REACTIONS IN HETEROGENEOUS SYSTEMS | 44
In the limit of 𝜂 → 1 (𝜙𝑛 → 0), there are no diffusion limits. In the limits of 𝜂 → 0 (for 𝜙𝑛 → ∞), there are
significant diffusion limits. For a first-order reaction in a sphere,
3(𝜙1 coth(𝜙1 ) − 1) 3
𝜂= 2 = 2 (𝜙𝑛 coth(𝜙𝑛 ) − 1)
𝑅 𝑘1 𝜌𝐵 𝑆𝐴 𝜙𝑛
𝐷𝐸𝐴
It turns out the right-hand form of this expression is true for other orders in spherical catalysts as well since
the reaction order only changes the power of 𝑛 − 1 in the Thiele modulus.
For a flat plate,
tanh(𝜙)
𝜂=
𝜙
The observed rate is
𝐸𝑎
−𝑟𝐴,obs = −𝜂𝑟𝐴,𝑠 = 𝜂𝑘0 𝑒 −𝑅𝑇 𝐶𝐴,𝑠
𝑛 𝑛
= 𝜂𝑘𝑛 𝐶𝐴,𝑠
where 𝑘𝑛 is the intrinsic rate constant. Also note that 𝜂𝑘𝑛 = 𝑘obs .
This can happen at small particle radius, small rate constants, high effective diffusivity, and/or low
temperatures. Therefore, when running at these conditions, one is measuring the real (i.e. intrinsic rate
constant).
𝑛
3 𝑛
3
𝜙 = large, − 𝑟𝐴,obs = 𝜂𝑘𝑛 𝐶𝐴,𝑠 = 𝑘𝑛 𝐶𝐴,𝑠 = √𝐷𝐸𝐴 √𝑘𝑛 𝐶𝐴,𝑠
𝜙𝑛 𝑅
REACTIONS IN HETEROGENEOUS SYSTEMS | 45
𝐸𝑎
Recall that 𝑘𝑛 is an intrinsic rate constant and can be expressed as √𝑘𝑛 = √𝑘𝑛,0 exp (− ) where 𝑘𝑛,0 is
2𝑅𝑇
the pre-exponential factor. This therefore says that for large 𝜙 (diffusion limitations), we have
𝐸𝐴
𝜙 = large, 𝐸𝐴,obs =
2
If one includes the temperature dependence of 𝐷𝐸𝐴 , it can be shown that
𝐸𝐴 + 𝐸diffusion
𝐸𝐴,obs =
2
The apparent order can be expressed as the following (for 𝑛 ≠ 1)
𝑛 − 1 𝑑 ln 𝜂
𝑛app = 𝑛 +
2 𝑑 ln 𝜙𝑛
such that the observed order is between the true order and 1.
One can find the diffusivity in the diffusion limiting regime by backing out 𝜂 from
𝑛
𝑟𝐴,obs = 𝜂𝑘𝑛 𝐶𝐴,𝑠
provided 𝑘𝑛 and 𝐶𝐴,𝑠 is the same between the two rates (e.g. if you change the catalyst size). If all the
parameters are the same except catalyst size, one can also state that
𝜙1 𝑅1
=
𝜙2 𝑅2
For the case of diffusion limitations, since 𝜂 ∝ 1/𝜙, one can state that
𝜂1 𝜙𝑛,2
=
𝜂2 𝜙𝑛,1
The above expression can be tested to see if one is operating in the diffusion-limiting regime. Conversely,
there are no diffusion limitations if changing the radius has no impact on the observed rate.
REACTIONS IN HETEROGENEOUS SYSTEMS | 46
𝑟𝐴,obs 𝑅2
𝜂𝜙𝑛2 =
𝐷𝐸𝐴 𝐶𝐴,𝑆
For 𝜙 ≪ 1, 𝜂 = 1 there is no pore diffusion limitations and so we expect 𝜂𝜙𝑛2 ≪ 1. For 𝜙 ≫ 1, 𝜂 ∝ 1/𝜙𝑛
there is strong pore diffusion limitations and 𝜂𝜙𝑛2 ≫ 1. Note that for first order, the above expression
simplifies to
𝑘obs 𝑅2
𝜂𝜙12 =
𝐷𝐸𝐴
Oftentimes, the above expression is written with a 𝑊 2 instead of 𝑅 2 such that for a sphere 𝑊 2 = (𝑅/3)2 .
This causes a difference of a factor of 9 but should not change the overall trend.
where 𝑘𝑔 is the mass transfer coefficient. Of course, the rate of reaction at the catalyst surface is given by
𝑛
𝑟𝐴 = 𝑘𝐶𝐴,𝑠 . Let us assume, for example, 𝑛 = 2. We can then equate these two expressions, solve for 𝐶𝐴,s ,
and plug back into 𝑟𝐴 to arrive at
𝑘𝑔 𝑘𝑔
𝑟𝐴 = 𝑘𝑔 ((1 + ) − √(1 + ) − 1) 𝐶𝐴,b
2𝑘𝐶𝐴,b 2𝑘𝐶𝐴,b
This is clearly neither 1st nor 2nd order but some intermediate. In the limit of 𝑘 ≫ 𝑘𝑔 , we arrive at
𝑟𝐴 = 𝑘𝑔 𝐶𝐴,s
which is 1st order and limited due to mass transfer. Contrastingly, in the limit of 𝑘𝑔 ≫ 𝑘, we arrive at
2
𝑟𝐴 = 𝑘𝐶𝐴,s
which is what we’d expect – 2nd order and a rate limited by reaction.
REACTIONS IN HETEROGENEOUS SYSTEMS | 47
𝑃𝐴
𝑃𝐴,𝑖
𝐶𝐴,𝑖
𝐶𝐴
𝐶𝐴,b
𝑦𝑔 𝑦𝐿
We will define the following gas film and liquid film mass transfer coefficients
REACTIONS IN HETEROGENEOUS SYSTEMS | 48
𝐷𝐴,𝑔 𝐷𝐴,𝐿
𝑘𝑔 = , 𝑘𝐿 =
𝑦𝑔 𝑦𝐿
𝑦 𝑦
𝐶𝐴,𝑖 sinh (𝛾 (1 − 𝑦 )) + 𝐶𝐴,b sinh (𝛾 𝑦 )
𝐿 𝐿
𝐶𝐴 =
sinh(𝛾)
where 𝛾 is the Hatta parameter defined as
𝑘 √𝑘𝐷𝐴,𝐿
𝛾 ≡ 𝑦𝐿 √ =
𝐷𝐴,𝐿 𝑘𝐿
where 𝑁𝐴,𝑖 is the flux at the interface, 𝐴𝑣 is the surface area per volume, and 𝑘𝐶𝐴,𝑖 is the reaction rate at the
interface. If we define the Sherwood number as
𝑘𝐿
Sh =
𝐴𝑣 𝐷𝐴
It can be shown that the effectiveness factor is
1 𝐶𝐴,𝑏 1
𝜂𝐿 = (1 − )
Sh γ tanh(𝛾) 𝐶𝐴,𝑖 cosh(𝛾)
1 𝐷𝐴
𝜂𝐿 = = 𝐴𝑣 √
𝛾 Sh 𝑘