You are on page 1of 41

Topic 3: Rate Laws and Stoichiometry

Rate Laws
Types of reactions (basic definitions)
• A homogeneous reaction is one that involves only one phase.

• A heterogeneous reaction involves more than one phase, and


reaction usually occurs at or very near the interface between the
phases.

• An irreversible reaction is one that proceeds in only one direction


and continues in that direction until the reactants are exhausted.
o A reversible reaction, on the other hand, can proceed in either
direction, depending on the concentrations of reactants and
products relative to the corresponding equilibrium concentrations.
o An irreversible reaction behaves as if no equilibrium condition
exists.
o Strictly speaking, no chemical reaction is completely irreversible
but, in many reactions, the equilibrium point lies so far to the right
that they are treated as irreversible reactions.
The Reaction Rate Constant
oIn the chemical reactions considered in the following
section, 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.

oThe limiting reactant is usually chosen as our basis


for calculation.
o The rate of dis-appearance of A, -r, depends on
temperature and composition.
o For many reactions it can be written as the product of a
reaction rate constant k and a function of the
concentrations (activities) of the various species involved
in the reaction:
The Reaction Rate Constant
• The algebraic equation that relates - r, to the species concentrations is
called the kinetic expression or rate law.

• The specific rate of reaction, kA, like the reaction rate always refers to
a particular species 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:
• We shall delete the subscript on the specific
reaction rate:
The Reaction Rate Constant
• The reaction rate constant k is not truly a constant, but is
merely independent of the concentrations of the species
involved in the reaction.

• The quantity k is also referred to as the specific reaction


rate.

• It is almost always strongly dependent on temperature.


o In gas-phase reactions, it depends on the catalyst and may be a
function of total pressure.

o In liquid systems it can also be a function of total pressure, and in


addition can depend on other parameters, such as ionic strength and
choice of solvent.

o These other variables normally exhibit much less effect on the specific
reaction rate than does temperature, so it will be assumed that k,
depends only on temperature.
The Reaction Rate Constant
• This assumption is valid in most laboratory and industrial
reactions and seems to work quite well.

• It was the great Swedish chemist Arrhenius who first


suggested that the temperature dependence of the specific
reaction rate, k, could be correlated by an equation of the
type:

(1)
• Where: A = pre-exponential factor or frequency factor, E =
activation energy, J/mol or cal/mol, R = gas constant = 8.3 14
J/mol. K = 1.987 cal/mol. K and T = absolute temperature, K.

• Equation (1), known as the Arrhenius equation, has been


verified empirically to give the temperature behaviour of most
reaction rate constants within experimental accuracy over
fairly large temperature ranges.
The Reaction Rate Constant
oThe activation energy E has been equated with a
minimum energy that must be possessed by
reacting molecules before the reaction will occur.

oFrom the kinetic theory of gases, the frequency


factor gives the fraction of the collisions between
molecules that together have this minimum
energy E.
Calculation of activation energy

• The activation energy is determined experimentally by carrying out the reaction at


several different temperatures. After taking the natural logarithm of Equation (2).

(2)

• it can be seen that a plot of versus 1/T should be a straight line whose slope is
proportional to the activation energy.

See Example 1 on the determination of the Activation Energy in the hand out.

• Calculate the activation energy for the decomposition of benzene diazonium


chloride to give chlorobenzene and nitrogen:
Elementary Rate Laws and Molecularity

• A reaction has an elementary rate law if the reaction order of each species is
identical with the stoichiometric coefficient of that species for the reaction.

• For example, the gas-phase reaction between hydrogen and iodine to form
hydrogen iodide:

• The rate law for this reaction is:

• In some circles when a reaction has an elementary rate law it is referred to as


an elementary reaction.
Molecularity

• The molecularity is the number of atoms, ions, or molecules involved


(colliding) in the rate-limiting step of the reaction.

• The terms unimolecular, bimolecular, and termolecular refer to reactions


involving, respectively, one, two, or 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:
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.

• That is, for the general reaction:

• The concentrations at equilibrium are related by the thermodynamic


relationship:
Reversible Reactions

Read more about reversible reactions in the hand out on chapter 3 on rate
laws and stoichiometry from page 88. Read also about activation energy and
look at the example 3-1 on how to calculate activation energy.
Stoichiometry
Stoichiometric Tables
• The stoichiometric table presents the stoichiometric relationships between
reacting molecules for a single reaction.

• That is, it tells us how many molecules of one species will be formed during a
chemical reaction when a given number of molecules of another species
disappears.

• These relationships will be developed for the general reaction:

(1)
Stoichiometry
• In formulating our stoichiometric table, we shall take species A as our
basis of calculation (i.e., limiting reactant) and then divide through by the
stoichiometric coefficient of A, in order to put everything on a basis of
“per mole of A.

(2)

• Next, we develop the stoichiometric relationship for reacting species that


give the change in the number of moles of each species (i.e., A, B, C, and
D).
Stoichiometry
Batch systems
• Figure 1 shows a batch system in which we will carry out the
reaction given by Equation (2).

• At time t = 0 we will open the reactor and place a number of


moles of species A, B, C, D, and I (NAo, NBO, NC0, ND0 and NI0
respectively) into the reactor.
• Species A is our basis of calculation and NA0 is the number of
moles of A initially present in the reactor.

• Of these, NA0X moles of A are consumed in the system as a

result of the chemical reaction, leaving (NAo - NAoX) moles of A


in the system. That is, the number of moles of A remaining in
the reactor after conversion X has been achieved is

NA = NA0 - NAOX = NAo (1 - X)


t=0

t=t
NA
NB
NC
ND
NI

Figure 1. Batch reactor


Stoichiometry
• The complete stoichiometric table for the reaction shown in Equation (2) taking
place in a batch reactor is presented in Table 1.

• To determine the number of moles of each species remaining after NA0X moles of
A have reacted, we form the stoichiometric table. This stoichiometric table
presents the following information:
• Column I: the particular species

• Column 2: the number of moles of each species initially present

• Column 3: the change in the number of moles brought about by reaction

• Column 4: the number of moles remaining in the system at time t


Stoichiometry
• For every mole of A that reacts, b/a moles of B must react;
therefore, the total number of moles of B that have reacted
is:
o Because B is disappearing from the system, the sign of the
"change” is negative.

o NB0 is the number of moles of B initially in the system.

o Therefore, the number of moles of B remaining in the

system, NB is given in the last column of the table above as:


o The stoichiometric coefficients in parentheses (d/a + c/a - b/a - 1)
represent the increase in the total number of moles per mole of A
reacted.

o The parameter tells us the change in the total number of moles per
mole of A reacted. The total number of moles can now be calculated
from the equation:
𝑁 𝑇 = 𝑁𝑇 0+ 𝛿 𝑁 𝐴0 𝑋
Reaction rate as a function of conversion x
• We recall from Chapter 1 that the kinetic rate law (e.g., -rA = kCA) is a function solely
of the intensive properties of the reacting materials (e.g., temperature, pressure,
concentration, and catalysts, if any).

• The reaction rate, -rA, usually depends on the concentration of the reacting species
raised to some power.

• Consequently, to determine the reaction rate as a function of conversion X, we


need to know the concentrations of the reacting species as a function of conversion.

𝑁
• The concentration of A is the number of moles of A per
𝐴 unit volume:
𝐶 𝐴 =
𝑉
• After writing similar equations for B, C, and D, we use the stoichiometric table to
express the concentration of each component in terms of the conversion X:
• We further simplify these equations by defining the
parameter , which allows us to factor NA0 in each of the
expressions, for concentration:
𝐶 𝐵=
𝑁 A0
( 𝑁 𝐵0
𝑁 𝐴0

𝑏
𝑎 ( ) )=
𝑋
(
𝑁 A 0 𝜃 𝐵− ( ) )
𝑏
𝑎
𝑋
𝛩 𝐵=
𝑁𝐵0
𝑁 𝐴0
𝑉 𝑉

𝐶𝐶=
𝑁 A0 ( 𝑁𝐶 0
𝑁 𝐴0
+
𝑐
𝑎 ( ) )=
𝑋 𝑁 A0
( 𝜃 𝐶+ ( ) 𝑋 ) (5 )𝛩 = 𝑁𝑁
𝑐
𝑎 𝐶
𝐶0

𝐴0
𝑉 𝑉

𝐶 𝐷=
𝑁 A0
( 𝑁 𝐷0
𝑁 𝐴0
+
𝑑
𝑎 ( ) )=
𝑋
(
𝑁 A 0 𝜃𝐷+ ( )𝑋)
𝑑
𝑎 𝛩 𝐷=
𝑁 𝐷0
𝑁 𝐴0
𝑉 𝑉
o We now need to only find volume as a function of conversion to obtain the
species concentration as a function of conversion.
Constant-Volume Batch Reaction Systems
• Some significant simplifications in the reactor design
equations are possible when the reacting system undergoes
no change in volume as the reaction progresses.

• These systems are called constant-volume, or constant-


density, because of the invariance of either volume or
density during the reaction process.
o This situation may arise from several causes.

o In gas-phase batch systems, the reactor is usually a sealed


vessel with appropriate instruments to measure pressure and
temperature within the reactor.
o The volume within this vessel is fixed and will not
change, and is therefore a constant-volume system.
o The laboratory bomb reactor is a typical example of this
type of reactor.
oAnother example of a constant-volume gas-phase
isothermal reaction occurs when the number of moles
of product equals the number of moles of reactant.

oThe water-gas shift reaction, important in coal


gasification and many other processes, is one of
these:
o In this reaction, 2 moles of reactant forms 2 moles of product.

o When the number of reactant molecules forms an equal


number of product molecules at the same temperature and
pressure, the volume of the reacting mixture will not change if
the conditions are such that the ideal gas law is applicable, or if
the compressibility factors of the products and reactants are
approximately equal.
o For liquid-phase reactions taking place in solution, the solvent
usually dominates the situation.
o As a result, changes in the density of the solute do not affect
the overall density of the solution significantly and therefore it
is essentially a constant-volume reaction process.
o Most liquid-phase organic reactions do not change density
during the reaction, and represent still another case to which
the constant-volume simplifications apply.
• An important exception to this general rule exists for
polymerization processes.
• For the constant-volume systems described above,
Equation (5) can be simplified to give the following
expressions relating concentration and conversion:
𝑁 A 0(1 − 𝑋 )
𝐶 𝐴= =𝐶 A 0 ( 1− 𝑋 )
𝑉0

𝑁 A0
( 𝜃𝐵 − ( ) 𝑋 ) =𝐶
𝑏

( ( ) 𝑋)
𝑎 𝑏
𝐶 𝐵= A0 𝜃𝐵 −
𝑉0 𝑎

(6)
𝑁 A0
( 𝜃 𝐶+
𝑐
( ) 𝑋)=𝐶
( ( ) 𝑋)
𝑎 𝑐
𝐶𝐶= A0 𝜃 𝐶+
𝑉0 𝑎

𝑁 A0
( ( ) )
𝜃𝐷+
𝑑
𝑋

( ( ) )
𝑎 𝑑
𝐶 𝐷= =𝐶 A0 𝜃𝐷+ 𝑋
𝑉0 𝑎
Look at examples 3.2 and 3.3 in Elements
Chemical Reaction Engineering by H. Scott
Fogler.

You might also like