Download as ppt, pdf, or txt
Download as ppt, pdf, or txt
You are on page 1of 16

Chemical Equilibrium

CHEMICAL EQUILIBRIUM - The state reached when the concentrations of reactants and
products remain constant over time. Chemical equilibrium exists when two opposing
reactions occur simultaneously at the same rate.

aA + bB  cC + dD
Chemical equilibria are dynamic equilibria; that is, individual molecules are continually
reacting, even though the overall composition of the reaction mixture does not change. In a
system at equilibrium, the equilibrium is said to lie toward the right if more C and D are
present than A and B, and to lie toward the left if more A and B are present.
Consider a case in which the coefficients in
the equation for a reaction are all 1. When
substances A and B react, the rate of the
forward reaction decreases as time passes
because the concentration of A and B
decrease.
At equilibrium the rates of both of these
processes are the same, so no net change takes
place over time:
forward rate = reverse rate
The equilibrium constant Kc, is defined as the product of the equilibrium concentrations
(moles per liter) of the products, each raised to the power that corresponds to its coefficient
in the balanced equation, divided by the product of the equilibrium concentrations of
reactants, each raised to the power that corresponds to its coefficient in the balanced
equation.
We have used the subscript “eq” to emphasize that the concentrations in the equilibrium
constant expression are those at equilibrium. For the remainder of this text, we shall omit
these subscripts, remembering that calculations with KC values always involve equilibrium
values of concentrations.
The thermodynamic definition of the equilibrium constant involves activities rather than
concentrations. The activity of a component of an ideal mixture is the ratio of its
concentration or partial pressure to a standard concentration (1 M) or pressure (1 atm). For
now, we can consider the activity of each species to be a dimensionless quantity whose
numerical value can be determined as follows.
1. For any pure liquid or pure solid, the activity is taken as 1.
2. For components of ideal solutions, the activity of each component is taken to be the ratio
of its molar concentration to a standard concentration of 1 M, so the units cancel.
3. For gases in an ideal mixture, the activity of each component is taken to be the ratio of its
partial pressure to a standard pressure of 1 atm, so again the units cancel.
Because of the use of activities, the equilibrium constant has no units; the values we put into
KC are numerically equal to molar concentrations, but are dimensionless, that is, they have
no units. In this text, calculations have usually included units along with numbers.
Calculations involving equilibrium are frequently carried out without units; we will follow
that practice in this text.
The magnitude of Kc is a measure of the extent to which reaction occurs. For any balanced
chemical equation, the value of Kc
1. is constant at a given temperature;
2. changes if the temperature changes;
3. does not depend on the initial concentrations.

A value of Kc much greater than 1 indicates that the “numerator concentrations” (products)
would be much greater than the “denominator concentrations” (reactants); this means that at
equilibrium most of the reactants would be converted into products.
If the substances involved in the reaction are all gases (as they are in the steam-reforming
and the water-gas shift reactions), then the equilibrium constant can also be expressed in
terms of partial pressures:


 i'
P   i'

KP is the equilibrium constant based on pressures products


KP 
P 
i
i

 i'
n   i'
reac tan ts
products
Kn 

 in   i Kn equilibrium constant based on numbers of moles
reac tan ts

 i'
x   i'

products
Kx 
KX (equilibrium constant with respect to mole fractions) 
 ix   i

reac tan ts
Relations between equilibrium constants

It is easier to measure the pressure of a gas than its concentration and, as long as the gas
behaves nearly ideally under the conditions of the experiment, the ideal gas law allows us to
relate these variables to each other:

  
 P i' x Ai' i' P  i'  x Ai' i' 
KP     K X P
 
 P i x Ai i P  i
 x Ai
i

Where  is the stoichiometric variation of reaction, i.e., the difference of stoichiometric


coefficients of products and of reactants.

Using ideal gas law, Pi V = ni R T (V – is here total volume occupied by the moles of i-th
component, ni), we may have a correlation between partial pressure and concentration:

ni
Pi  R T  ci R T
v
Also, using the Dalton’s law for mixtures Pi = P xi , the mole fraction is substituted by its
definition so:
ni
Pi  P
 ni

we now shall obtain correlations of KP with both KC and Kn

 i'
 P 
   ni' i' 
 ci'  R T 
 i' i  P 
  ni' 
KP   KC   R T  
  K n  
 ci  R T 
i i 
 P  i   ni 
   ni i
  ni 


 P 
KP  KX P 
 KC  RT  
 K n  
  ni 
Factors that affect equilibria. Le Chatelier’s Principle
We can "perturb" chemical reactions at equilibrium by adding more of a reactant or product.
For example, increasing the concentration of a product causes the reaction to run in reverse,
converting some of the added product to reactants. This response to perturbation, or stress, is
described in a principle named after French chemist Henri Louis Le Chatelier (1850-1936).
Le Chatelier's principle states that if a system at equilibrium is subjected to a stress, the
position of the equilibrium shifts in the direction that relieves that stress.

Three types of changes can disturb the equilibrium of a reaction.


1. Changes in concentration
2. Changes in pressure or volume (for reactions that involve gases)
3. Changes in temperature

1.Changes in Concentration

When a system at equilibrium is disturbed by a change in concentration of one of the


components, the system reacts in the direction that reduces the change:
• If the concentration increases, the system reacts to consume some of it.
• If the concentration decreases, the system reacts to produce some of it.
As an example of how a change in concentration affects an equilibrium, let’s consider the
reaction in aqueous solution of iron(III) and thiocyanate (SCN -) ions to give an equilibrium
mixture that contains the Fe–N bonded red complex ion FeNCS2+

Shifts in the position of this equilibrium can be detected by observing how the color of the
solution changes when we add various reagents.

Color changes produced by adding various reagents to an equilibrium mixture of Fe 3+(pale


yellow), SCN-(colorless), and FeNCS2+(red): (a) The original solution. (b) After adding FeCl3
to the original solution, the red color is darker because of an increase in FeNCS 2+ ,(c) After
adding KSCN to the original solution, the red color again deepens. (d) After adding H2C2O4
-oxalic acid (a poisonous substance present in plants such as rhubarb) to the original solution,
the red color disappears because of a decrease in FeNCS 2+ the yellow color is due to
Fe(C2O4)33-; (e) After adding HgCl2 to the original solution, the red color again vanishes.
Example:
The reaction of iron(III) oxide with carbon monoxide occurs in a blast furnace when
iron ore is reduced to iron metal:

Fe2O3(s) + 3CO(g)  2Fe(s) + 3CO2(g)

Use Le Châtelier’s principle to predict the direction of net reaction when an equilibrium
mixture is disturbed by:
(a) Adding Fe2O3;
(b) Removing CO2;
(c) Removing CO.

SOLUTION
(a) Because Fe2O3 is a solid, its “concentration” doesn’t change when more Fe 2O3 is added.
Therefore, there is no concentration stress, and the original equilibrium is undisturbed.
(b) Le Châtelier’s principle predicts that the concentration stress of removed CO 2 will be
relieved by net reaction from left to right to replenish the CO 2.
(c) Le Châtelier’s principle predicts that the concentration stress of removed CO will be
relieved by net reaction from right to left to replenish the CO.
2.Changes in Pressure and Volume
Changes in pressure have significant effects only on equilibrium systems with gaseous
components. Aside from phase changes, a change in pressure has a negligible effect on
liquids and solids because they are nearly incompressible. Pressure changes can occur in
three ways:
• Changing the concentration of a gaseous component
• Adding an inert gas (one that does not take part in the reaction)
• Changing the volume of the reaction vessel
We just considered the effect of changing the concentration of a component, and that
reasoning holds here. Next, let's see why adding an inert gas has no effect on the equilibrium
position. Adding an inert gas does not change the volume, so all reactant and product
concentrations remain the same. In other words, the volume and the number of moles of the
reactant and product gases do not change, so their partial pressures do not change.
In general, for reactions that involve gaseous reactants or products, Le Chatelier’s
Principle allows us to predict the following results.
1. If there is no change in the total number of moles of gases in a reaction, a volume
(pressure) change does not affect the position of equilibrium.
2. If a reaction involves a change in the total number of moles of gases:
(a) A decrease in volume (increase in pressure) shifts a reaction in the direction that
produces the smaller total number of moles of gas.
(b) An increase in volume (decrease in pressure) shifts a reaction in the direction that
produces the larger total number of moles of gas.
3.Changes in Temperature

Of the three types of disturbances-a change in concentration, in pressure, or in temperature-


only temperature changes alter K. To see why, we must take the heat of reaction into
account:
PCl3(g) + Cl2(g)  PCl5(g)

The forward reaction is exothermic (releases heat; ΔH < 0), so the reverse reaction is
endothermic (absorbs heat; ΔH > 0):
PCl3(g) + Cl2(g)  PCl5(g) + heat (exothermic)
PCl3(g) + Cl2(g)  PCl5(g) + heat (endothermic)

If we consider heat as a component of the equilibrium system, a rise in temperature means


heat is "added" to the system and a drop in temperature means heat is "removed" from the
system. As with a change in any other component, the system shifts to reduce the effect of
the change. Therefore, a temperature increase (adding heat) favors the endothermic (heat-
absorbing) direction, and a temperature decrease (removing heat) favors the exothermic
(heat-releasing) direction.

Thus,
• A temperature rise will increase KC for a system with a positive ΔH.
• A temperature rise will decrease KC for a system with a negative ΔH.
4. Addition of a Catalyst

Adding a catalyst to a system changes the rate of the reaction, but this cannot shift the
equilibrium in favor of either products or reactants. Because a catalyst affects the
activation energy of both forward and reverse reactions equally, it changes both rate
constants by the same factor, so their ratio, KC, does not change. Adding a catalyst to a
reaction at equilibrium has no effect; it changes KC.
The same equilibrium mixture is achieved with or without the catalyst, but the
equilibrium is established more quickly in the presence of a catalyst. Not all reactions
attain equilibrium; they may occur too slowly, or else products or reactants may be
continually added or removed. Such is the case with most reactions in biological
systems. On the other hand, some reactions, such as typical acid–base neutralizations,
achieve equilibrium very rapidly.
Laws of chemical equilibrium (van’t Hoff law)
The reaction isotherm (van’t Hoff, 1886)
 r G   r G 0  R T ln QP
QP term is exactly similar in form to the expression of equilibrium constant KP; however,
the partial pressures for various participating substances do not correspond to the conditions of
equilibrium. The expression for the free energy of a reaction consists of two parts:
(1) a constant term whose value depends only on the reaction taking place, and
(2) a variable term that depends on the partial pressures of the reactants and the
products, the stoichiometry of the reaction, and the temperature.
The equation is a first expression of the reaction isotherm (the van’t Hoff isotherm,
1886), in conditions of non-equilibrium. It is applied, for instance, in the starting of a chemical
reaction, by knowing the initial partial pressures of all participants (existing species both as
reactants and products).
For a reacting system situated exactly in an equilibrium position, there is no net
change of the Gibbs free energy, because the free energy of the forward reaction is exactly
balances that of the backward reaction. Consequently, ΔrG = 0, provided the temperature and
pressure are not allowed to alter.
  P  i' 
0   r G 0  R T ln  
Ai'
  PA  i 
 i  at equilibrium

 r G 0   R T ln K P
The equilibrium constant of a reaction may therefore be calculated from standard free
energy data and vice versa.

The equation represents the mathematical form of the reaction isotherm in conditions of
chemical equilibrium. Remind that ΔrG0 must depend upon the particular choice of standard
state; its tabulated values in the majority of books refer for the state of 1 atm pressure,
although recently this state is accepted as 1 bar for all gaseous species as participants.
Calculation of the equilibrium composition
Homogeneous gaseous equilibria
In the following some schemes are shown to calculate the equilibrium constant by
using the extent of reaction, ξ, introduced by De Donder equation: ni = ni0 + υiξ .
For chemical reactions without change in the number of moles (without number of
moles variation), Δυ = 0, a unique equilibrium constant K will be involved:
Some examples are as follows:
K P = K X = KC = K n
H2 + I2 ↔
2 HI
initial state: a b
0
consumed: ξ ξ
0
at equilibrium: a-ξ b-ξ

Kn 
n 2
HI

 2 
2

4 2
nH 2 nI 2  a    b     a    b   

You might also like