You are on page 1of 48

Chemical Reaction Kinetics

Lecture#2

Dr. Laith S. Sabri


Junior
The Rate of Reaction, –rA
• The rate of reaction tells us how fast a number of moles of one chemical
species are being consumed to form another chemical species.
• The term chemical species refers to any chemical component or element
with a given identity. The identity of a chemical species is determined by
the kind, number, and configuration of that species’ atoms.
• For example, the species para-xylene is made up
of a fixed number of specific atoms in a definite molecular arrangement or
configuration.
• The structure shown illustrates the kind, number, and configuration of
atoms on a molecular level. Even though two chemical compounds
have exactly the same kind and number of atoms of each element, they
could still be different species because of different configurations.
• For example, 2-butene has four carbon atoms and eight hydrogen atoms;
however, the atoms in this compound can form two different
arrangements.
The molecularity of a reaction is the number of atoms, ions, or molecules
involved (colliding) in a reaction step. The terms unimolecular, bimolecular,
and termolecular refer to reactions involving, respectively, one, two, and
three atoms (or molecules) interacting or colliding in any one reaction step.
The most common example of a unimolecular reaction is radioactive decay,
such as:
the spontaneous emission of an alpha particle from uranium-238 to give
thorium and helium

The rate of disappearance of uranium (U) is given by the rate law

The only true bimolecular reactions are those that involve the collision with free radicals
(i.e., unpaired electrons, e.g., Br•), such as:

with the rate of disappearance of bromine given by the rate law


The probability of a termolecular reaction, where three molecules collide all at once, is
almost nonexistent, and in most instances the reaction pathway follows a series of
bimolecular reactions, as in the case of the reaction,

The reaction pathway for this “Hall of Fame” reaction is quite interesting and
is discussed in Chapter 9, along with similar reactions that form active intermediate
complexes in their reaction pathways.
In discussing Rates of Chemical Reaction, we have three rates to
consider:
• Relative Rates
• Rate Laws
• Net Rates
Relative Rates tell us how fast one species is disappearing or appearing
relative to the other species in the given reaction.

Rate Laws are the algebraic equations that apply to a given reaction.

Net Rates of formation of a given species (e.g., A) is the sum of the rate of reactions of A all
the reactions in which A is either a reactant or product in the system.

A thorough discussion of net rates is given in Chapter 8 where we discuss multiple reactions.
Reaction rate is measured by the amount of product produced or reactants consumed per
unit time,

• [B] concentration of products will increase over time.


• [A] concentration of reactants will decrease over time

The Progress of a Simple


Reaction (A → B). The
mixture initially contains only
A molecules (purple). Over
time, the number of A
molecules decreases and more
B molecules (green) are
formed (top). The graph
shows the change in the
number of A and B molecules
in the reaction as a function of
time over a 1 min period
(bottom).
https://chem.libretexts.org/Bookshelves/General_Chemistry/Map%3A_Chemistry_-_The_Central_Science_(Brown_et_al.)/14%3A_Chemical_Kinetics/14.2%3A_Reaction_Rates
Rate of reaction rA is:

Rate of reaction per unit weight of catalyst • a function of concentration,


and rate of reaction per unit volume is • temperature,
related through the bulk density of the • pressure, and
catalyst particle in the fluid media • the type of catalyst (if any).

Independent of the type of reaction


system (batch, plug flow, etc.) but An algebraic equation, not a
dependent on the reaction chemistry. differential equation.

Factors Affecting Reaction Rates


1- Concentration of the reactants. When the reactant concentration increases, the reaction
rate increases.
2-Temperature:
• Faster molecules collide more often and collisions have more energy.
• Most reactions, even exothermic reactions, require energy to occur.

3- Pressure, When the gaseous reactant pressure increases, more reactant is compressed into a
given volume (i.e. the reactant concentration reactant concentration increases), the reaction
rate increases.
4- Presence of Catalysts,
- increase rates of chemical reactions without being used up.
- Rate-accelerating agents.
- Not consumed in the reaction.

5. Chemical nature of the reactants,

-What are the chemical the reactants?


‐Do the reactants have strong bonds or weak bonds?
‐What about the number of bonds that need to be broken and reformed?

6. Physical State of the Reactants,


-Surface area consideration, The greater the surface area, the greater the ability the
reactants can meet, and therefore, the greater the reaction rate.
-Phase consideration ,Homogeneous or Heterogeneous reaction
Rate of Reaction =

Rate of Reaction r =

Always with respect to a given reactant or product


• r=[reactants] decrease with time (disappear).
• r=[products] increase with time (formation).

Rate of Reaction =

rB = rate of formation of B(product) per unit volume

-rA = rate of a disappearance of A(reactant) per unit volume

Note: Always positive whether something is disappear or formation in [X ]


Conversion and yield:
Conversion, X, is defined as the fraction (or percentage) of the more important or limiting
reactant that is consumed. With two reactants A and B and a nearly Stoichiometric feed,
conversions based on each reactant could be calculated.

Yield, Y, is the amount of desired product produced relative to the amount that would have
been formed if there were no byproducts and the main reaction went to completion.
Relative Rates of Reaction
The relative rates of reaction of the various species involved in a reaction can be obtained from
the ratio of the stoichiometric coefficients. For Reaction below,
+ → +
we see that for every mole of A that is consumed, c/a moles of C appear. In other words,

Similarly, the relationship between the rates of formation of C and D is

-1 mole of A and of B consumed , while mole of C and

mole of D formed or appear

Rate of reaction or disappearance of A = -rA


Then the reaction Stoichiometry ;
EXAMPLE,

Answer:
we have

the rate of disappearance of NO is

and the rate of disappearance of oxygen, O2, is


The Reaction Order and the Rate Law
In the chemical reactions considered in the following paragraphs, we take as the basis
of calculation a species A, which is one of the reactants that is disappearing as a result
of the reaction. The limiting reactant is usually chosen as our basis for calculation.
The rate of disappearance of A, –rA, depends on temperature and concentration. For
many irreversible reactions, it can be written as the product of a reaction rate
constant, kA , and a function of the concentrations (activities) of the various species
involved in the reaction:

Concept 1. Law of Mass Action.


The rate of reaction increases with increasing concentration of reactants
owing the corresponding increase in the number of molecular collisions.
The rate law for bimolecular collisions is derived in the collision theory
section of the Professional Reference Shelf R3.1. A schematic of the
reaction of A and B molecules colliding and reacting is shown in Figure
R3.1.
R3.1. Collision Theory
In this section, the fundamentals of collision theory
The rate law gives the relationship between reaction rate
and concentration.
The algebraic equation that relates –rA to the species concentrations is called the
kinetic expression or rate law. The specific rate of reaction (also called the rate
constant), kA , like the reaction rate, –rA , always refers to a particular species in the
reaction and normally should be subscripted with respect to that species. However,
for reactions in which the stoichiometric coefficient is 1 for all species involved in
the reaction, for example,
Power Law Models and Elementary Rate Laws
The dependence of the reaction rate, –rA, on the concentrations of the species present, fn(Cj), is
almost without exception determined by experimental observation. Although the functional
dependence on concentration may be postulated from theory, experiments are necessary to
confirm the proposed form. One of the most common general forms of this dependence is the
power law model. Here the rate law is the product of concentrations of the individual reacting
species, each of which is raised to a power, for example,

The exponents of the concentrations in Equation above lead to the concept of reaction order.
The order of a reaction refers to the powers to which the concentrations are raised in the
kinetic rate law. In Equation above, the reaction is α order with respect to reactant A, and β
order with respect to reactant B. The overall order of the reaction, n, is

Overall reaction order


The units of –rA are always in terms of concentration per unit time, while the units of the
specific reaction rate, kA, will vary with the order of the reaction. Consider a reaction
involving only one reactant, such as
with an overall reaction order n. The units of rate, –rA, and the specific reaction rate, k are:

Generally, the overall reaction order can be deduced from the units of the specific
reaction rate constant k. For example, the rate laws corresponding to a zero-, first-,
second-, and third-order reaction, together with typical units for the corresponding
rate constants, are
An elementary reaction is one that involves a single reaction step, such as the bimolecular
reaction between an oxygen free radical and methanol molecule

The stoichiometric coefficients in this reaction are identical to the powers in the rate law.
Consequently, the rate law for the disappearance of molecular oxygen is

The reaction is first order in an oxygen free radical and first order in methanol; therefore, we
say that both the reaction and the rate law are elementary.
For example, the oxidation reaction of nitric oxide discussed earlier
is not really an elementary reaction, but follows an
elementary rate law; therefore,
Note: The rate constant, k, is defined with respect to NO.

Where do you find rate laws?


The rate constants and the reaction orders for a large number of gas- and
liquid-phase reactions can be found in the National Bureau of Standards’ circulars
and supplements.
- Kinetic data for a larger number of reactions can be obtained on CD-ROMs provided by
National Institute of Standards and Technology (NIST).

Additional sources are


Tables of Chemical Kinetics: Homogeneous Reactions, National Bureau of Standards
Circular 510.
Nonelementary Rate Laws
A large number of both homogeneous and heterogeneous reactions do not follow simple rate
laws. Examples of reactions that don’t follow simple elementary rate laws are discussed below.

• Do not follow law of mass action


• proceeds in more than one step
• Involves the appearance of intermediates
Intermediates cannot be observed as they are highly reactive and
present in minute quantities , intermediates could be:
Homogeneous Reactions
The overall order of a reaction does not have to be an integer, nor does the order have to be an
integer with respect to any individual component. As an example, consider the gas-phase
synthesis of phosgene,

in which the kinetic rate law is

This reaction is first order with respect to carbon monoxide, three-halves order with respect to
chlorine, and five-halves order overall. Sometimes reactions have complex rate expressions
that cannot be separated into solely temperature-dependent and concentration-dependent
portions. In the decomposition of nitrous oxide,

the kinetic rate law is

Both and are strongly temperature dependent. When a rate expression such as the
one given above occurs, we cannot state an overall reaction order. Here, we can only speak of
reaction orders under certain limiting conditions.
For example, at very low concentrations of oxygen, the second term in the denominator would
be negligible with respect to 1 (1 >> ), and the reaction would be “apparent” first order
with respect to nitrous oxide and first order overall. However, if the concentration of oxygen
were large enough so that the number 1 in the denominator were insignificant in comparison
with the second term, ( >> 1), the apparent reaction order would be –1 with respect
to oxygen and first order with respect to nitrous oxide, giving an overall apparent zero order.
Rate expressions of this type are very common for liquid and gaseous reactions promoted by
solid catalysts (see Chapter 10 in *). They also occur in homogeneous reaction systems with
reactive intermediates (see Chapter 9 in *).
It is interesting to note that although the reaction orders often correspond to the stoichiometric
coefficients, as evidenced for the reaction between hydrogen and iodine, just discussed to form
HI, the rate expression for the reaction between hydrogen and another halogen, bromine, is
quite complex. This nonelementary reaction,

proceeds by a free-radical mechanism, and its reaction rate law is

*H. Scott Fogler ”Elements of Chemical Reaction Engineering”. 5th edition 2016
Heterogeneous Reactions.
Historically, it has been the practice in many gas-solid catalyzed reactions to write the rate law
in terms of partial pressures rather than concentrations. In heterogeneous catalysis it is the
weight of catalyst that is important, rather than the reactor volume. Consequently, we use − in
order to write the rate law in terms of mol per kg of catalyst per time in order to design PBRs.
An example of a heterogeneous reaction and corresponding rate law is the hydrodemethylation
of toluene (T) to form benzene (B) and methane (M) carried out over a solid catalyst

The rate of disappearance of toluene per mass of catalyst, − , i.e., (mol/mass/time) follows
Langmuir-Hinshelwood kinetics (discussed in Chapter 10), and the rate law was found
experimentally to be

where the prime − in notes typical units are in per kilogram of catalyst (mol/kg-cat/s), PT, ,
and PB are partial pressures of toluene, hydrogen, and benzene in (kPa or atm), and KB and KT
are the adsorption constants for benzene and toluene respectively, with units of kPa–1 (or atm–1).
The specific reaction rate k has units of
You will find that almost all heterogeneous catalytic reactions will have a term such as
in the denominator of the rate law (cf. Chapter 10)
To express the rate of reaction in terms of concentration rather than partial pressure, we
simply substitute for Pi using the ideal gas law

The rate of reaction per unit weight (i.e., mass) catalyst − , (e.g., − ), and the rate of
reaction per unit volume, − , are related through the bulk density ρb (mass of solid/volume) of
the catalyst particles in the fluid media:
Relating rate per unit volume
and rate by per unit mass of
catalyst
In fluidized catalytic beds, the bulk density, , is normally a function of the
volumetric flow rate through the bed.
Consequently, using the above equations for Pi and − we can write the rate law for the
hydromethylation of toluene in terms of concentration and in (mole/dm3) and the rate, − in
terms of reactor volume, i.e.,
In summary on reaction orders, they cannot
be deduced from reaction stoichiometry.
Even though a number of reactions follow
elementary rate laws, at least as many
reactions do not. One must determine the
reaction order from the literature or from
experiments.
Reversible Reactions
All rate laws for reversible reactions must reduce to the thermodynamic relationship relating
the reacting species concentrations at equilibrium. At equilibrium, the rate of reaction is
identically zero for all species (i.e., -rA ≡ 0). That is, for the general reaction,
the concentrations at equilibrium are related by the
thermodynamic relationship for the equilibrium
constant KC
The units of the thermodynamic
equilibrium constant, KC, are

To illustrate how to write rate laws for reversible reactions, we will use the combination of
two benzene molecules to form one molecule of hydrogen and one of diphenyl. In this
discussion, we shall consider this gas-phase reaction to be elementary and reversible

or, symbolically,

The forward and reverse specific reaction rate constants, kB and k-B, respectively, will be
defined with respect to benzene. Benzene (B) is being depleted by the forward reaction
in which the rate of disappearance of benzene is
If we multiply both sides of this equation by -1, we obtain the expression for the rate of formation
of benzene for the forward reaction
(1)
For the reverse reaction between diphenyl (D) and hydrogen (H2),

the rate of formation of benzene is given as


(2)

Again, both the rate constants kB and k-B are defined with respect to benzene!!!
The net rate of formation of benzene is the sum of the rates of formation from the forward
reaction [i.e., Equation (3)] and the reverse reaction [i.e., Equation (2)]

(3)
Multiplying both sides of Equation (3) by -1, and then factoring out kB, we obtain the rate law
for the rate of disappearance of benzene, -rB

Replacing the ratio of the reverse to forward rate law constants by the
reciprocal of the concentration equilibrium constant, KC, we obtain

(4)

where

The equilibrium constant decreases with increasing temperature for exothermic reactions and
increases with increasing temperature for endothermic reactions.
Let’s write the rate of formation of diphenyl, rD, in terms of the concentrations of hydrogen,
H2, diphenyl, D, and benzene, B. The rate of formation of diphenyl, rD, must have the same
functional dependence on the reacting species concentrations as does the rate of
disappearance of benzene, –rB. The rate of formation of diphenyl is

(5)
Using the relationship given by Equation (6) for the general reaction

(6)

we can obtain the relationship between the various specific reaction rates, kB, kD

Comparing Equations (4) and (5), we see the relationship


between the specific reaction rate with respect to diphenyl, kD,
and the specific reaction rate with respect to benzene, kB, is
Consequently, we see the need to define the rate constant, k, with respect
to a particular species. Finally, we need to check to see if the rate law
given by Equation (3-14) is thermodynamically consistent at equilibrium.
Applying Equation (3-10) (and Appendix C) to the diphenyl reaction and
substituting the appropriate species concentration and exponents,
thermodynamics tells us that
Now let’s look at the rate law. At equilibrium, –rB ≡0, and the rate law given by Equation (3-
14) becomes
At equilibrium, the rate law must
Reduce to an equation consistent wth
Thermodynamic equilibrium.
Rearranging, we obtain, as expected, the
equilibrium expression

that is identical to Equation (3-17) obtained from thermodynamics. From Appendix C,


Equation (C-9), we know that when there is no change in the total number of moles and the
heat capacity term, CP = 0, the temperature dependence of the concentration equilibrium
constant is

(C-9)

Therefore, if we know the equilibrium constant at one


temperature, T1 [i.e., KC (T1)], and the heat of reaction, ∆ ,
we can calculate the equilibrium constant at any other
temperature T. For endothermic reactions, the equilibrium
constant, KC, increases with increasing temperature; for
exothermic reactions, KC decreases with increasing
temperature. A further discussion of the equilibrium constant
and its thermodynamic relationship is given in Appendix C.
For large values of the equilibrium constant, KC, the reaction
behaves as if it were irreversible.
Rates and the Reaction Rate Constant
There are three molecular concepts relating to the rate of reaction we will discuss:
Concept 1. Law of Mass Action. The rate of reaction increases with increasing
concentration of reactants owing to the increased number of molecular collisions at the
higher reactant concentrations. Concept 1: rate laws, and the dependence of the rate
reactant concentration.
We now discuss:
Concept 2, potential energy surfaces and energy barriers, and
Concept 3, the energy needed for crossing the barriers.

The Rate Constant k


The reaction rate constant k is not truly a constant; it is merely independent of the
concentrations of the species involved in the reaction. The quantity k is referred to as either the
specific reaction rate or the rate constant. It is almost always strongly dependent on
temperature. It also depends on whether or not a catalyst is present, and in gas-phase reactions,
it may be a function of total pressure. In liquid systems it can also be a function of other
parameters, such as ionic strength and choice of solvent. These other variables normally exhibit
much less effect on the specific reaction rate than does temperature, with the exception of
supercritical solvents, such as supercritical water. Consequently, for the purposes of the material
presented here, it will be assumed that kA depends only on temperature. This assumption is valid
in most laboratory and industrial reactions, and seems to work quite well.
It was the great Nobel Prize–winning Swedish chemist Svante Arrhenius
(1859–1927) who first suggested that the temperature dependence of the specific
reaction rate, kA, could be correlated by an equation of the type

Svante August Arrhenius ( 19 February 1859 – 2


October 1927) was a Swedish scientist. Originally a
physicist, but often referred to as a chemist,
Arrhenius was one of the founders of the science of
physical chemistry. He received the Nobel Prize for
Chemistry in 1903, becoming the first Swedish
Nobel laureate. In 1905, he became director of the
Nobel Institute, where he remained until his death.
Arrhenius was the first to use basic principles of
physical chemistry to estimate the extent to which
increases in atmospheric carbon dioxide are
responsible for the Earth's increasing surface
temperature. In the 1960s, David Keeling https://en.wikipedia.org/wiki/Svante_Arrhenius

demonstrated that human-caused carbon dioxide


emissions are large enough to cause global
warming.
Arrhenius equation

(3-18)

where
A = pre-exponential factor or frequency factor
E = activation energy, J/mol or cal/mol
R = gas constant = 8.314 J/mol • K = 1.987 cal/mol • K
T = absolute temperature, K

Equation (3-18), known as the Arrhenius equation, has been


verified empirically to give the correct temperature behavior of
most reaction rate constants within experimental accuracy over
fairly large temperature ranges. The Arrhenius equation is
derived in the Professional Reference Shelf R3.A: Collision
Theory on the CRE Web site.

Additionally, one can view the activation energy in terms of collision theory (Professional
Reference Shelf R3.1). By increasing the temperature, we increase the kinetic energy of the
reactant molecules. This kinetic energy can in turn be transferred through molecular collisions
to internal energy to increase the stretching and bending of the bonds, causing them to reach
an activated state, vulnerable to bond breaking and reaction.
Why is there an activation energy? If the reactants are free radicals that essentially react
immediately on collision, there usually isn’t an activation energy. However, for most atoms and
molecules undergoing reaction, there is an activation energy. A couple of the reasons are that in
order to react
1. The molecules need energy to distort or stretch their bonds so that
they break and now can form new bonds.
2. The molecules need energy to overcome the steric and electron repulsive
forces as they come close together.
The activation energy can be thought of as a barrier to energy transfer (from kinetic energy to
potential energy) between reacting molecules that must be overcome. The activation energy is
the minimum increase in potential energy of the reactants that must be provided to transform the
reactants into products. This increase can be provided by the kinetic energy of the colliding
molecules.
In addition to the concentrations of the reacting species, there are two
other factors that affect the rate of reaction,
• the height of the barrier, i.e., activation energy, and
• the fraction of molecular collisions that have sufficient energy to
cross over the barrier (i.e., react when the molecules collide).
If we have a small barrier height, the molecules colliding will need only low kinetic energies to
cross over the barrier. For reactions of molecules with small barrier heights occurring at room
temperatures, a greater fraction of molecules will have this energy at low temperatures.
However, for larger barrier heights, we require higher temperatures where a higher fraction of
colliding molecules will have the necessary energy to cross over the barrier and react. We will
discuss each of these concepts separately.
Concept 2. Potential Energy Surfaces and Energy Barriers.
One way to view the barrier to a reaction is through the use of potential energy surfaces and the
reaction coordinates. These coordinates denote the minimum potential energy of the system as
a function of the progress along the reaction path as we go from reactants to an intermediate to
products. For the exothermic reaction

the potential energy surface and the reaction coordinate are shown in Figures 3-1 and 3-2. Here
EA, EC, EAB, and EBC are the potential energy surface energies of the reactants A, BC and
product molecules (AB and C), and EABC is the energy of the complex A–B–C at the top of
the barrier. Figure 3-1(a) shows the 3–D plot of the potential energy surface, which is
analogous to a mountain pass where we start out in a valley and then climb up to pass over the
top of the pass, i.e., the col or saddle point, and proceed down into the next valley. Figure 3-
1(b) shows a contour plot of the pass and valleys and the reaction coordinate as we pass over
the col from valley to valley.

Figure 3-1 A potential energy surface


for the H + CH3OH H2 + CH2OH
from the calculations of Blowers and
Masel. The lines in the figure are
contours of constant energy. The lines
are spaced 5 kcal/mol. Richard I.
Masel, Chemical Kinetics and
Catalysis, p. 370, Fig.7.6 (Wiley,
2001).
Figure 3-2 Progress along reaction path. (a)
Symbolic reaction; (b) Calculated from
computational software on the CRE Web site,
Chapter 3 Web Module. (c) Side view at point
X in Figure 3-1(b) showing the valley.
The Arrhenius Plot
Postulation of the Arrhenius equation, Equation (3-18), remains the greatest single advancement
in chemical kinetics, and retains its usefulness today, more than a century later. The activation
energy, E, is determined experimentally by measuring the reaction rate at several different
temperatures. After taking the natural logarithm of Equation (3-18), we obtain

(3-24)

We see that the activation energy can be found from a plot of ln kA as a function of (1/T), which
is called an Arrhenius plot. The larger the activation energy, the more temperature-sensitive the
reaction. That is, for large E, an increase in just a few degrees in temperature can greatly
increase k and thus increase the rate of reaction.

Calculation of the
activation energy

Examples in pages 91

Figure 3-6 Calculation of the activation


energy from an Arrhenius plot
Activation energy is a measure of the minimum energy that the reacting molecules must
have in order for the reaction to occur (energy required to reach transition state). At the
same concentration but different two temperature Activation Energy can be estimated as :

Or

This equation says that if we know the specific reaction rate k(T1) at a temperature, T1,
and we know the activation energy, E, we can find the specific reaction rate k(T2) at
any other temperature, T2, for that reaction.

Heat
Absorbed
Heat
Released

Activation energy for exothermic and endothermic reaction


Heat of Reaction
Information on process
– Feeds (e.g. A , B) are reactants
– Outlets are products (e.g. C, D) of the reaction
– Heat is released or absorbed during the process
– Exothermic reaction: ΔHr is negative
– Endothermic reaction: ΔHr is positive
Stoichiometric reaction,
Where |vi| is stoichiometric
coefficient
Where the value of vi is positive for products and negative for reactants.

Heat of Reaction at temperature T and is designated ΔHr(T) [is always given per mole of the
species that is the basis of calculation, i.e., species A [Joules per mole of A reacted]

The molal enthalpy of species i at a particular temperature and pressure, Hi , is usually


expressed in terms of an enthalpy of formation of species i at some reference temperature TR,
, plus the change in enthalpy, , that results when the temperature is raised from the reference
temperature, TR, to some temperature T
The reference temperature at which ( ) is given is usually at TR =25o C. For any substance
i that is being heated from T1 to T2 in the absence of phase change
for specific heat
varied with
temperature

for constant specific heat with


Temperature throughout

substitute for the enthalpy of each species, we have

Heat of reaction at the reference temperature Over all change in the Heat capacity per mole
TE (heat of formation) ∆ ( ) a reacted ∆
How Temperature dependency of Rate of Reaction?
• Which reactions are more temperature sensitive (high E or low E)??
• high E large temperature sensitive
• low E small temperature sensitive
• Effect of temperature on exothermic & endothermic reversible reactions??
• Exothermic reaction: EForward < Ebackward
• Endothermic reaction: EForward > EBackward
• Are reactions sensitive at low temperature ranges or high temperature ranges??
• reactions are more sensitive at low temperature range for doubling rate of reaction

shows temperature dependency of the reaction rate shows temperature dependency of the
and rate constant reaction rate
Van't Hoff Equation
The van't Hoff equation relates the change in the equilibrium constant, Keq, of a chemical
reaction to the change in temperature, T, given the standard enthalpy change, ΔHo, for the
process.
The van't Hoff equation may be derived from the Gibbs free energy and van’t Hoff’s isotherm.
Gibbs free energy (G) when the reaction occurs at constant temperature and pressure. For a
system at constant T & P. Gibbs free energy can be written as:

Where H is enthalpy(kJmol-1), S is entropy(Jmol-1K-1) and T is temperature (K) At any given


temperature, the change in Gibbs free energy:-

if we impose the standard state (note that standard state does not specify a temperature!) on all
species present, then we get

the relation between equilibrium constant and standard reaction


Gibbs energy is so-called van’t Hoff’s isotherm , which can be
obtained as :-
For a reaction to be spontaneous, the overall free energy at any
At equilibrium, ΔG=0, then equation 21 ,

equating the two expressions equation 22 & 20 for ΔG∘, then

You might also like