Download as pdf or txt
Download as pdf or txt
You are on page 1of 17

CHEMICAL EQUILIBRIUM 1

CHEMICAL EQUILIBRIUM
1. Irreversible and Reversible Reactions
Those reactions which proceeds in forward direction and reaches almost to completion are called
irreversible reactions. For example

2KClO3(s) 
MnO 2
→ 2KCl(s) + 3O 2(g) (Thermal decomposition)

NaOH (aq ) + HCl(aq) 


→ NaCl(aq) + H 2 O (Strong acid–strong base neutralisation reaction)
2AgNO3 + BaCl 2 
→ 2AgCl ↓ + Ba(NO3 )2 (Precipitation reaction)
Whereas, those reactions which proceed in forward and backward directions both and never
reaches completion are called reversible reactions. These reactions can be initiated in any
direction. For example,

PCl5(g) PCl3(g) + Cl 2(g) (Thermal dissociation)

CH 3COOH + NaOH CH3 COONa + H 2 O (Neutralisation reactions)


Weak acid strong base

N 2(g ) + 3H 2(g ) 2NH3(g) (synthesis reaction)


But when Fe(s) is heated and water vapour is passed over it in open vessel, it is converted to
Fe3O4(s) along with the evolution of hydrogen gas.
3Fe(s) + 4H 2 O(g) 
→ 3Fe3O 4(s) + 4H 2(g)
and when Fe3O4 is reduced with hydrogen gas, it gives Fe(s) and H2O
Fe3O 4(s) + 4H 2(g) 
→ 3Fe (s) + 4H 2 O (g)
But, if the reaction is carried out in a closed container, this reaction becomes reversible.
3Fe(s) + 4H 2O (g) Fe3O 4(s) + 4H 2(g)

2. State of Equilibrium

It has generally been observed that many changes (physical and chemical) do not proceed to
completion when they are carried out in a closed container. Consider for example vapourisation
of water,
Water Vapour
At any temperature, vapourisation of water takes place, initially the concentration of water is
much greater than the concentration of vapour, but with the progress of time, concentration of
vapour increases whereas that of water remains constant and after a certain interval of time,
there is no change in concentration of vapour, this state is known as state of physical equilibrium.
In a similar way, this has also been found for chemical reactions, for example, when PCl5(g) is
heated in a closed container, its dissociation starts with the formation of PCl3(g) and Cl2(g). Initially,
only PCl5(g) was taken, but with the progress of reaction, PCl3(g) and Cl2(g) are formed due to
dissociation of PCl5(g). After a certain interval of time, the concentration of PCl5(g), PCl3(g) and
2 CHEMICAL EQUILIBRIUM

Cl2(g) each becomes constant. It does not mean that at this point of time, dissociation of PCl5(g)
and its formation from PCl3(g) and Cl2(g) has been stopped. Actually the rate of dissociation of
PCl5 and the rate of formation of PCl5(g) becomes equal. This state is called the state of chemical
equilibrium. So, the state of chemical equilibrium is dynamic.
PCl 5(g) PCl3(g) + Cl2(g)
This can be shown graphically.

Equilibrium
Forward
reaction

Rate of reaction
Concentration

State of equilibrium
Backward
reaction
Reactant

Time Time
So, “state of chemical equilibrium can be defined as the state when the rate of forward reaction
becomes equal to rate of reverse reaction and the concentration of all the species becomes
constant”.

3. Law of Mass Action

Guldberg and Waage in 1807 gave this law and according to this law, “At constant temperature,
the rate at which a substance reacts is directly proportional to its active mass and the rate at
which a chemical reaction proceeds is directly proportional to the product of active masses of
the reacting species”.
The term active mass of a reacting species is the effective concentration or its activity (a) which
is related to the molar concentration (C) as
a = f ⋅C
where f = activity coefficient. f < 1 but f increases with dilution and as V → ∞ i.e. C → 0,
f →1 i.e., a → C. Thus at very low concentration the active mass is essentially the same as
the molar concentration. It is generally expressed by enclosing the formula of the reacting species
in a square bracket.
To illustrate the law of mass action, consider the following general reaction at constant
temperature,
aA (g) + bB(g) cC(g) + dD(g)
Applying law of mass action,
Rate of forward reaction,
Rf ∝ [A]a [B]b
or Rf = kf[A]a[B]b …(i)
Where kf = rate constant of forward reaction
Similarly,
CHEMICAL EQUILIBRIUM 3
Rate of reverse reaction
Rr ∝ [C]c [D]d
or Rr = kr[C]c [D]d …(ii)
where kr = rate constant of reverse reaction.
At equilibrium,
Rate of forward reaction = Rate of reverse reaction
i.e., Rf = Rr
So, from equations (i) and (ii), we get
kf[A]a [B]b = kr[C]c[D]d
k f [C]c [D]d
or, =
k r [A]a [B]b

kf [C]c [D]d
or = Kc =
kr [A]a [B]b
Where, Kc is the equilibrium constant in terms of molar concentration.
Equilibrium Constant (Kp) in terms of Partial Pressures: Consider the same
general reaction taking place at constant temperature,
aA(g) + bB(g) cC(g) + dD(g)
From law of mass action,
[C]c [D]d
Kc = …(1)
[A]a [B]b
From ideal gas equation
PV = nRT
n
or, P= RT
V
At constant temperature,
n
P∝
V
Thus, we can say in a mixture of gases, Partial pressure of any component (say A)
PA ∝ [A]
Similarly, PB ∝ [B]
PC ∝ [C]
PD ∝ [D]
So, equation (1) can be rewritten as
PCc × PDd
Kp = …(2)
PAa × PBb

4. Relationship Between Kp and Kc

From the above, for the same general reaction at constant temperature.
[C]c [D]d
Kc = …(1)
[A]a [B]b
4 CHEMICAL EQUILIBRIUM

PCc × PDd
and, Kp = …(2)
PAa × PBb
From the ideal gas equation,
PV = nRT
n
P= RT
V
So, PB = [B]RT ; PD = [D]RT
Similarly, PA = [A]RT; PC = [C]RT
Substituting the values of PA, PB, PC and PD in equation (2), we get

[C]c [D]d (RT)(c+ d)


or Kp = ×
[A]a [B]b (RT)(a + b)

or K p = K c ⋅ (RT) ∆n
Where, ∆n = (c + d) – (a + b)
i.e., ∆n = sum of no. of moles of gaseous products – sum of no. of moles of gaseous reactants.
If ∆n > 0, Kp > Kc
If ∆n = 0, Kp = Kc
and if ∆n < 0, Kp < Kc
Illustration 1:
Calculate the Kc and Kp for the following reactions and also deduce the relationship
between Kc and Kp
1 3
(i) 2SO2(g) + O2(g) 2SO3(g) (ii) N 2(g) + H 2(g) NH 3(g)
2 2
Solution:
(i) 2SO2(g) + O2(g) 2SO3(g)
Applying law of mass action,
[SO3 ]2
Kc = …(i)
[SO 2 ]2 [O 2 ]
2
PSO
Kp = 3
…(ii)
2
PSO 2
× PO2
We know that
K p = K c (RT)∆n
Kc
∆n = 2 – (2+1) = – 1 ∴ K p =
RT
1 3
(ii) N 2(g) + H 2(g) NH 3(g)
2 2
[NH3 ]
Kc = …(i)
[N 2 ]1/ 2 [H 2 ]3/ 2
CHEMICAL EQUILIBRIUM 5
PNH3
Kp = …(ii)
PN1/22 × PH3/2 2

Now, K p = K c (RT)∆n
3 1
∆n = 1 −  +  = −1
2 2
K
Kp = c
RT

Illustration 2:
At 400K for the gaseous reaction
2A + 3B 3C + 4D
the value of Kp is 0.05. Calculate the value of Kc (R = 0.082 dm3 atm K–1 mol–1)
Solution:
From the given data
For the reaction 2A + 3B 3C + 4D
∆n = (3 + 4) – (2 +3) = 2
Since R is given in dm3 atm. K–1 mol–1, we shall take P = 1 atm.
Substituting these values in the equation
∆n 2
 CRT   1 × 0.082 × 400 
K p = Kc   , we get 0.05 = K c  
 P   1 
0.05
Kc =
0.082 × 0.082 × 400 × 400
Kc = 4.648 × 10–5

Illustration 3:
(i) The vapour density of a mixture consisting of NO2 and N2O4 is 38.3 at 26.7°C.
Calculate the number of moles of NO2 in 100 gm mixture.
(ii) Establish a relationship between Kc and Kp for the following reactions.
(a) N2(g) + O 2(g) 2NO (g)

()b) (NH 4 )2 CO3(s) 2NH 3(g) + CO2(g) + H 2O(g)


Solution:
D−d
(i) N2O4 2NO2 α=
d
46 − 38.3
t=0 1.087 0 α= = .2 (20%)
38.3
t = teq 1.087 (1 – α) 1.087 × 2α
∴ n NO = 1.087 × 0.4 = 0.435
2

(ii) (a) ∆ng = 0 ∴ Kp = Kc


(b) ∆ng = 4 ∴ Kp = Kc (RT)4
6 CHEMICAL EQUILIBRIUM

5. Application of Law of Mass Action

1. Synthesis of Hydrogen Iodide:


Suppose ‘a’ moles of H2 and ‘b’ moles of I2 are heated at 444°C in a closed container of
volume ‘V’ litre and at equilibrium, 2x moles of HI are formed.
H2(g) + I2(g) 2HI(g)
a b
Initial concentration (mol L–1) 0
V V
a−x
Equilibrium concentration(mol L–1)
V
[HI]2
Kc = …(i)
[H 2 ][I 2 ]
Substituting the equilibrium concentrations of H2, I2 and HI in equation (i), we get
2
 2x 
  4x 2
Kc =  V =
 a − x  b − x  (a − x)(b − x) …(ii)
  
 V  V 
2
PHI
Kp = …(iii)
PH 2 × PI2
Suppose the total pressure of the equilibrium mixture at 444°C is P, then
PHI = mole fraction of HI × Total pressure
2x
PHI = ×P
a+b
Similarly,
(a − x) b−x
PH2 = × P and PI2 = ×P
(a + b) (a + b)
Substituting these values in equation (iii) we get
2
 2x  2
  P
Kp = a+b
a−x b−x
 ×P× × P
a+b a+b
4x 2
∴ Kp = …(iv)
(a − x)(b − x)
So, one can see from equations (iii) and (iv), that
Kp = Kc
This is so, because ∆n= 0 for the synthesis of HI from H2 and I2.
2. Thermal Dissociation of Phosphorus Pentachloride
PCl5(g) dissociates thermally according to the reaction,
PCl5(g) PCl3(g) + Cl2(g)
Let us consider that 1 mole of PCl5 has been taken in a container of volume V litre and at
equilibrium x moles of PCl5(g) dissociates. Thus
CHEMICAL EQUILIBRIUM 7

PCl5(g) PCl3(g) + Cl2(g)


Initial concentration (mol L–1) 1 0 0
1− x x
Equilibrium concentration(mol L–1)
V V
According to law of mass action, at constant temperature,
[PCl3 ][Cl 2 ]
Kc = [PCl5 ] …(1)

PPCl3 × PCl2
and Kp = PPCl5 …(2)

Substituing the values of equilibrium concentration, in equation (1), we have

x2
Kc = or Kc =
V(1 − x)

Now, total number of moles at equilibrium = 1 – x + x + x = 1 + x


x
Mole fraction of PCl3 = Mole fraction of Cl2 =
1+ x
1− x
and mole fraction of PCl5 =
1+ x
Suppose total pressure at equilibrium is P, then we have from equation (2),
x x 2
× P
x x
× Kp = 1 + x 1 + x Kp =
x2P
V V 1− x or
P (1 − x 2 )
1− x 1+ x
V Similarly, we can apply law of mass action on any reaction at equilibrium.

Illustration 4:
Derive an expression for Kc and Kp for the reaction
N2(g) + 3H2(g) 2NH3(g)
Assuming that in a container of volume V, initially 1 mole of N2 and 3 moles of H2 were
taken and at equilibrium 2x moles of NH3 is formed.
Solution:
N2 + 3H2 2NH3
t=0 1 3 –
t = teq 1–x 3 – 3x 2x = 4 – 2x
(2x)2 x 2V 2
∴ Kc = 2
×V =
(1 − x) (3 − 3x)3 27(1 − x ) 4

(2x)2 (4 − 2 x) 2 16 x 2 (2 − x ) 2
Kp = × =
(1 − x) (3 − 3x)3 P2 27(1 − x ) 4 P 2
8 CHEMICAL EQUILIBRIUM

Illustration 5:
The degree of dissociation at a certain or given temperature of PCl5 at 2 atm is found
to be 0.4. At what pressure, the degree of dissociation of PCl5 will be 0.6 at the same
temperature? Also calculate the equilibrium constant for the reverse reaction.
Solution:
The dissociation of PCl5 takes place according to the equation,
PCl5 PCl3(g) + Cl2(g)
Let 1 mole of PCl5 is taken in a closed container. Then
PCl5(g) PCl3(g) + Cl2(g)
Mole(s) before dissociation 1 0 0
Mole(s) at equilibrium 1 – α α α
As degree of dissociation(α) is given at 2 atm
1 – α = 1.0 – 0.4 = 0.6
∴ Total moles at equilibrium = 1 – α + α + α = 1 – 0.4 + 0.4 + 0.4 = 1.4
0.4 0.4
PPCl3 × PCl2 × 2× ×2
Kp = 1.4 1.4
Now, = 0.6
PPCl5 ×2
1.4

0.4 × 0.4 × 4 × 1.4 0.32


Kp = =
0.6 × 2 × 1.4 × 1.4 0.84
Now, as temperature remains the same, Kp will also remain the same. So for
α = 0.6
PCl5(g) PCl3(g) + Cl2(g)
Mole before dissociation 1 0 0
Moles at equilibrium 1 – 0.6 0.6 0.6
0.6 0.6
× × P2 0.6 × 0.6 × 1.6
Again, K p = 1.6 1.6 ⇒ Kp = P
0.4 0.4 × 1.6 × 1.6
×P
1.6
32 9 32 × 4 128
= P ∴ P= = = 0.677 atm
84 16 9 × 21 189

Illustration 6:
(i) Calculate the percentage dissociation of H2S(g), if 0.1 mole of H2S is kept in
0.4 litre vessel at 1000K. For the reaction, 2H2S(g) 2H2(g) + S2(g), the
value of Kc = 1.6 × 10–6
(ii) A sample of HI was found to be 22% dissociated when equilibrium was reached.
What will be the degree of dissociation if hydrogen is added in the proportion
of 1 mole for every mole of HI originally present, the temperature and volume
of the system being kept constant?
Solution:
(i) 2H2 + S2(g)
2H2S(g) Kc = 1.6 × 10–6
t=0 0.1 – –
t = teq 0.1 – x
x x/2
value of Kc suggest that (0.1 – x) ~
= 0.1
CHEMICAL EQUILIBRIUM 9

x3 1
∴ ×
2 0.4
= 1.6 × 10–6 ⇒ x = 2.34 ×10–3
2(0.1)
2.34 ×10 −3
∴ α= × 100 = 2.34%
0.1
(ii) 2HI H2 + I2
t=0 1 – –
t = teq 1– α α/2 α/2 since α = 0.22
(0.11) 2 1
∴ Kc = 2 = 64
(0.88)
now 2HI H2 + I2
t=0 1 1 0
t = teq 1– α α/2 α/2

1 (1 + α / 2).(α / 2) 2α + α 2
= =
64 (1 − α ) 2 4(1 − α ) 2
∴ 4 + 4α2 – 8α = 128α + 64α2
∴ 60α2 + 136α – 4 = 0
∴ α = 0.29

6. Equilibrium Constant for Heterogeneous Equilibria


The equilibrium which involves reactants and products in different physical states. The law of
mass action can also be applied on heterogeneous equilibria as it was applied for homogeneous
equilbria (involving reactants and products in same physical states).
(i) Thermal Dissociation of Solid Ammonium Chloride:
The thermal dissociation of NH4Cl(s) takes place in a closed container according to the equation:
NH4Cl(s) NH3(g) + HCl(g)
Let us consider 1 mole of NH4Cl(s) is kept in a closed container of volume ‘V’ litre at
temperature TK and if x mole of NH4Cl dissociates at equilibrium, then
NH4Cl(s) NH3(g) + HCl(g)
Initial moles 1 0 0
Moles at equilibrium 1 – x x x
[NH ][HCl]
Applying law of mass action, K c = [NH Cl]
3

As NH4Cl is a pure solid, so there is no appreciable change in its concentration. Thus,


Kc = [NH3] [HCl]
x x x2
Kc = × =
V V V2
and K p = PNH3 × PHCl

(ii) Thermal Dissociation of Ag2CO3:


Ag2CO3(s) dissociates thermally according to the equation.
Ag2CO3(s) Ag2O(s) + CO2(g)
Applying law of mass action, at constant temperature, we get,
10 CHEMICAL EQUILIBRIUM

[Ag 2O][CO 2 ]
Kc =
[Ag 2 CO 3 ]
Now, let us consider that 1 mol of Ag2CO3(s) is heated in a closed container of volume V and
x mol of Ag2CO3(s) dissociates at equilibrium, then
Ag2CO3(s) Ag2O(s) + CO2(g)
Initial moles 1 0 0
Equilibrium moles 1 – x x x
Now, as Ag2CO3 and Ag2O are solids, so their concentration can be assumed to be constants
thus
x
K c = [CO 2 ] =
V
K P = PCO2

This principle, which is based on the fundamentals of a stable equilibrium, states that
“When a chemical reaction at equilibrium is subjected to any stress, then the
equilibrium shifts in that direction in which the effect of the stress is reduced”.
Confused with “stress”. Well by stress here what we mean is any change of reaction conditions
e.g. in temperature, pressure, concentration etc.
This statement will be explained by the following example.
endo
Let us consider the reaction: 2NH3 (g) exo
N2 (g) + 3H2 (g)
Let the moles of N2, H2 and NH3 at equilibrium be a, b and c moles respectively. Since the
reaction is at equilibrium,
( ) (X )( X )
3 3
PN 2 × PH2 N 2 .PT H 2 .PT
= Kp =
(P ) (X )
2 2
NH3 NH 3 .PT

Where,
X terms denote respective mole fractions and PT is the total pressure of the system.
3
 a   b 
 × PT   × PT 
a+b+c  a+b+c  =K
⇒ 2 p
 c 
 × PT 
a+b+c 

a
Here, = mole fraction of N2
a +b+c
b
= mole fraction of H2
a +b+c
c
= mole fraction NH3
a +b+c
ab3
×
( PT ) 2

⇒ 2 = Kp
(a + b + c)
2
c

( a + b + c ) RT
Since PT = ( assuming all gases to be ideal)
V
CHEMICAL EQUILIBRIUM 11
2
ab3  RT 
∴ ×  = KP …(1)
c2  V 
Now, let us examine the effect of change in certain parameters such as number of moles,
pressure, temperature etc.
If we increase a or b, the left hand side expression becomes QP ( as it is disturbed from
equilibrium) and we can see that QP > KP
The reaction therefore moves backward to make QP = KP.
If we increase c, Q P < K P and the reaction has to move forward to revert
back to equilibrium.
If we increase the volume of the container (which amounts to decreasing the pressure), QP <
KP and the reaction moves forward to attain equilibrium.
If we increase the pressure of the reaction, then equilibrium shifts towards backward direction
since in reactant side we have got 2 moles and on product side we have got 4 moles. So
pressure is reduced in backward direction.
If temperature is increased, the equilibrium will shift in forward direction since the forward
reaction is endothermic and temperature is reduced in this direction.
However from the expression if we increase the temperature of the reaction, the left hand side
increases (QP) and therefore does it mean that the reaction goes backward (since QP > KP)?.
Does this also mean that if the number of moles of reactant and product gases are equal, no
change in the reaction is observed on the changing temperature (as T would not exist on the
left hand side)?. The answer to these questions is No. This is because KP also changes with
temperature. Therefore, we need to know the effect of temperature on both QP and KP to
decide the course of the reaction.
Effect of Addition of Inert Gases to A Reaction At Equilibrium
1. Addition at constant pressure
Let us take a general reaction
aA + bB cC + dD
c d
 n   n 
 C × PT   D × PT 
We know, Kp = 

 n   n
  ∑ 

a b
 n   n 
 A
×P   B
× PT 
 ∑
 n T 
  ∑
n 

Where,
nC nD, nA, nB denotes the no. of moles of respective components and PT is the total pressure
and ∑n = total no. of moles of reactants and products.
∆n
nc × nd  P 
Now, rearranging , K P = ac Db ×  T 

n A × n B  n 
Where ∆n = (c + d) – (a + b)
Now, ∆n can be = 0, < 0 or > 0
Let us take each case separately.
(a) ∆n = 0 : No effect (b) ∆n = ‘+ve’ :
∆n
 P   P 
Addition of inert gas increases the ∑n i.e.  T  is decreased and so is  T  . So

 n ∑ 
 n  ∑
products have to increase and reactants have to decrease to maintain constancy of Kp.
So the equilibrium moves forward.
12 CHEMICAL EQUILIBRIUM
c) ∆n = ‘–ve’ :
∆n
 P   P 
In this case  n  decreases but  
T T

 ∑   n
 ∑ increases. So products have to decrease and
reactants have to increase to maintain constancy of Kp. So the equilibrium moves backward.
2. Addition at Constant Volume :
Since at constant volume, the pressure increases with addition of inert gas and at the same
∆n
 P 
time ∑n also increases, they almost counter balance each other. So  n 
T
can be safely
  ∑
approximated as constant. Thus addition of inert gas has no effect at constant volume.
Dependence of Kp or Kc on Temperature
Now we will derive the dependence of KP on temperature.
Starting with Arrhenius equation of rate constant
− Ea f / RT
kf = Af e … (i)
Where, kf = rate constant for forward reaction, Af = Arrhenius constant of forward reaction,
E af = Energy of activation of forward reaction
− E a r / RT
k r = Are …(ii)
Dividing (i) by (ii) we get,
( E ar − E af )
kf Af Ea f
= e RT a

k r Ar E ar E
Energy

k
We know that k = K (equilibrium constant )
f

( E ar − E af ) ∆H
k A
∴ K = f = fe RT
k r Ar Reaction co-ordinate

At temperature T1
( E a r − Ea f )
A RT1
K T1 = f e …(iii)
Ar
At temperature T2
( Ea r −Ea f )
A RT2
K T2 = fe … (iv)
Ar
Dividing (iv) by (iii) we get
 Ea r − E a f   1 1
K T2    −  K
T2 ar E −E  1
af 1
⇒ log K = 2.303 R  T − T 
 R   T2 T1 
=e
K T1 T1  2 1
CHEMICAL EQUILIBRIUM 13

The enthalpy of a reaction is defined in terms of activation energies as E af − E a r = ∆H


K T2 −∆H  1 1
∴ log =  − 
K T1 2.303 R  T2 T1 

K T2 ∆H 1 1 
∴ log K = 2.303 R  T - T  …(v)
T1  1 2

For an exothermic reaction, ∆H would be negative. If we increase the temperature of the system
(T2>T1), the right hand side of the equation (V) becomes negative.
∴ K T2 < K T1 , that is, the equilibrium constant at the higher temperature would be less than
that at the lower temperature.
Now let us analyse our question. Will the reaction go forward or backward?
Before answering this, we must first encounter another problem. If temperature is increased,
the new KP would either increase or decrease or may remain same. Let us assume it increases.
Now, QP can also increase, decrease or remain unchanged. If KP increases and QP decreases,
then Q PT2 < K PT2 , therefore the reaction moves forward. If KP increase and QP remains same,
then . Again, the reaction moves forward. What, if KP increase and QP also
increases?
Will QPT2 = K PT2 or or ? This can be answered by simply looking at
the dependence of QP and KP on temperature. You can see from the equation that KP depends
on temperature exponentially. While Q’s dependence on T would be either to the power
Q
2
( )
Q PPTT2 >
<QK PT < K
PT 2
1
PT
2
g,l,t…….. Therefore the variation in KP due to T would be more than in QP due to T.
∴ KP would still be greater than QP and the reaction moves forward again.
Therefore, to see the temperature effect, we need to look at KP only. If it increases, the reaction
moves forward, if it decreases, reaction moves backward and if it remains fixed, then no change
at all.

Illustration 7:
Under what conditions will the following reactions go in the forward direction ?
1. N2(g)+ 3H2(g) 2NH3(g) + 23 k cal.
2. N2(g) + O2(g) 2NO(g) - 43.2 k cal.
3. C(s) + H2O(g) CO2(g) + H2(g) + X k cal.
4. N2O4(g) 2NO2(g) - 14 k cal.
Solution:
1. Low T, High P, excess of N2 and H2.
2. High T, any P, excess of N2 and O2
3. Low T, Low P, excess of C and H2O
4. High T, Low P, excess of N2O4.

Illustration 8:
Under what conditions will the following reactions go in the forward direction?
1. 2SO2(g) + O2(g) 2SO3(g) + 45 k cal.
2. 2NO(g) + O2(g) 2NO2(g) + 27.8 k cal.
3. PCl5(g) PCl3(g) + Cl2(g)- X k cal.
14 CHEMICAL EQUILIBRIUM
Solution :
(i) Low T, high P, excess of O2 and SO2
(ii) Low T, high P, excess of NO and O2
(ii) High T, low P, excess of PCl5

Le Chatelier’s principle, as already stated, is applicable to all types of equilibria involving not
only chemical but physical changes as well. A few examples of its application to physical equilbria
are discussed below.
1. Vapour pressure of a liquid: Consider the equilibrium
Liquid Vapour
It is well known that the change of a liquid into its vapour is accompanied by absorption of
heat whereas the conversion of vapour into liquid state is accompanied by evolution of heat.
According to Le Chatelier’s principle, therefore, addition of heat to such a system will shift
the equilibrium towards the right. On raising the temperature of the system, liquid will evaporate.
This will raise the vapour pressure of the system. Thus, the vapour pressure of a liquid increases
with rise in temperature.
2. Effect of pressure on the boiling point of a liquid: The conversion of liquid
into vapour, as represented by the above equilibrium, is accompanied by increase of pressure
(vapour pressure). Therefore, if pressure on the system is increased, some of the vapours will
change into liquid so as to lower the pressure. Thus, the application of pressure on the system
tends to condense the vapour into liquid state at a given temperature. In order to counteract
it, a higher temperature is needed. This explains the rise of boiling point of a liquid on the
application of pressure.
3. Effect of pressure on the freezing point of a liquid (or melting point of
a solid): At the melting point, solid and liquid are in equilibrium:
Solid Liquid
Now, when a solid melts, there is usually a change, either increase or decrease, of volume.
For example, when ice melts, there is decrease in volume, or at constant volume, there is
decrease in pressure. Thus, increase of pressure on ice water system at a constant
temperature will cause the equilibrium to shift towards the right, i.e., it will cause the ice to
melt. Hence, in order to retain ice in equilibrium with water at the higher pressure it will be
necessary to lower the temperature. Thus, the application of pressure will lower the melting
point of ice.
When sulphur melts, there is increase in volume or at constant volume, there is increase in
pressure. From similar considerations, it follows that if the pressure on the system, sulphur (solid)
sulphur (liquid) is increased, the melting point is raised.
4. Effect of temperature on solubility: In most cases, when a solute passes into solution,
heat is absorbed, i.e., cooling results. Therefore according to Le Chatelier’s principle, when
heat is applied to a saturated solution in contact with solute, the change will take place in that
direction which absorbed heat (i.e., which tends to produce cooling). Therefore, some more
of the solute will dissolve. In other words, the solubility of the substance increases with rise in
temperature.
Dissociation of a few salts (e.g., calcium salts of organic acids) is accompanied by evolution
of heat. In such cases, evidently, the solubility decreases with rise in temperature.
CHEMICAL EQUILIBRIUM 15

The Gibbs free energy function is a true measure of chemical affinity under conditions of constant
temperature and pressure. The free energy change in a chemical reaction can be
defined as
∆G = G(products) – G(reactants)
When ∆G = 0, there is no net work obtainable. The system is in a state of equilibrium. When
∆G is positive, net work must be put into the system to effect the reaction, otherwise it cannot
take place. When ∆G is negative, the reaction can proceed spontaneously with accomplishment
of the net work. The larger the amount of this work that can be accomplished, the farther
away is the reaction from equilibrium. For this reason –∆G has often been called the driving
force of the reaction. From the statement of the equilibrium law, it is evident that the driving
force depends on the concentration of the reactants and products. It also depends upon the
temperature and pressure which determine the molar free energies of the reactants and products.
The reaction conducted at constant temperature (i.e., in a thermostat)
–∆G = – ∆H + T∆S
The driving force is made up of two parts, –∆H term and T∆S term. The –∆H term is the
heat of reaction at constant pressure and T∆S is heat involved when the process is carried
out reversibly. The difference is the amount of heat of reaction which can be converted into
net work (–∆G), i.e., total heat minus unavailable heat.
If the reaction is carried out at constant volume, the decrease in Helmholtz function –∆G = –
∆E + T∆S would be the proper measure of affinity of the reactant or the driving force of the
reaction.
Now we can see why Berthollet and Thompson were wrong in assuming that driving force of
the reaction was the heat of reaction. They neglected the T∆S term. The reasons for the
apparent validity of their principle was that for many reactions, ∆H term far outweighs the T∆S
term. This is especially true at low temperature, since at higher temperature, T∆S term increases.
The fact that driving force for a reaction is large (∆G is large negative quantity) does not mean
that the reaction will necessarily occur under any given conditions.
For example, the reaction
1
H 2 + O 2 
→ H 2 O; ∆G = −228.6kJ
2
does not occur at the laboratory temperature. The reaction mixture may be kept for years
without any detectable formation of water. Here ∆H factor favours, but ∆S factor disfavours
the reaction.
Similarly, the reaction
2Mg + O2  2MgO; ∆G = – 570.6kJ
is not favoured. However, the thermite reaction
3
2Al + O ? Al2O3
2 2
with large value of – ∆G proceeds favourably.
Standard Free Energy and Equilibrium Constant: The change in free energy
for a reaction taking place between gaseous reactants and products represented by the general
equation.
aA + bB cC + dD
According to Van’t Hoff reaction isotherm
pcC × pdD
∆G = ∆G 0 + RT ln = ∆G0 + RTlnQp
paA × p bB
16 CHEMICAL EQUILIBRIUM
the condition for a system to be at equilibrium is that
∆G = 0
Thus at equilibrium
 p c × pd 
0 = ∆G 0 + RT ln  aC Db  = ∆G 0 + RT ln K p0
 p A × pB 
Whence ∆G0 = – RTlnK0p
−∆G 0
Hence ln K 0p =
RT
Note: 1. In the reaction, where all gaseous reactants and products; K represents Kp
2. In the reaction, where all solution reactants and products; K represents Kc
3. a mixture of solution and gaseous reactants; Kx represents the thermodynamic equilibrium
constant and we do not make the distinction between Kp and Kc.
we may conclude that for standard reactions, i.e., at 1 M or 1 atm.
When ∆G0 = –ve or K > 1: forward reaction is feasible
∆G0 = +ve or K < 1: reverse reaction is feasible
∆G0 = 0 or K = 1: reaction is at equilibrium (very rare)
Illustration 9:
Kc for the reaction N2O4 2NO2 in chloroform at 291 K is 1.14. Calculate the
free energy change of the reaction when the concentration of the two gases are 0.5
mol dm–3 each at the same temperature. (R = 0.082 lit atm K–1 mol–1.)
Solution:
From the given data
T = 291 K; R = 0.082 lit atm K–1 mol–1
Kc = 1.14; C NO2 = C N2 O4 = 0.5 mol dm–3
The reaction quotient Qc for the reaction
[NO 2 ]2 0.5 × 0.5
Qc = = = 0.5
[N 2 O 4 ] 0.5
Since Qp = Qc(RT)∆n and ∆n = 2 – 1 = 1 in this case
∴Qc = 0.5 (0.082 × 291) = 11.93
Kp = Kc(RT)∆n = 1.14 (0.082 × 291) = 27.1
Substituting these values in the equation
∆G = ∆G0 + RTlnQp = – RT ln Kp + RT ln Qp, we get
= – 2.303 RT (log Kp – log Qp)
∆G = – (0.082 × 291 × 2.303) [log 27.2 – log 11.93]
= – 54.95 (1.4346 – 1.0766) = – 19.67 lit atm

Illustration 10:
Calculate the pressure of CO2 gas at 700K in the heterogeneous equilibrium reaction
CaCO3(s) CaO(s) + CO2(g) if ∆G0 for this reaction is 130.2 kJ mol–1.
Solution:
Here Kp = PCO2
Also, ∆G 0 = −RT ln K p

∆G 0 130.2 × 103 Jmol−1


∴ ln K p = = ∴ pCO2 = K p = 1.94 × 10−10 atm
RT (8.314JK −1mol−1 )(700K)
CHEMICAL EQUILIBRIUM 17
Illustration 11:
For the equilibrium NiO (s) + CO (g) Ni (s) + CO 2(g) , ∆ G 0 (J mol –1 )
= – 20,700 – 11.97 T. Calculate the temperature at which the product gases at
equilibrium at 1 atm will contain 400 ppm (parts per million) of carbon monoxide.
Solution: For the given reaction K p = pCO2 / p CO
Since

∆G 0 = − RT ln K p

∆G 0 20,700 + 11.97T
∴ ln K p = =
RT RT
The equation when solved for T using R = 8.314 K–1 mol–1, gives T = 399K.

Illustration 12:
i) Calculate the partial pressure of HCl gas above solid a sample of NH4Cl(s) as a result of
its decomposition according to the reaction:
NH4Cl(s) NH 3(g) + HCl (g) K p = 1.04 × 10–16
ii) Calculate the equilibrium constant of a reaction at 300 K if ∆G0 at this temperature
for the reaction is 29.29 kJ mol–1.
Solution :
(i) NH4Cl(s) NH3(g) + HCl(g) Kp = 1.04 ×10–16
p <<1pCO2 , hence
1
p =
KCO = = 2,500
Kp = PNH × PHCl ⇒ P2 = Kp ∴ P = Kp
pCO 400 × 10−6 3

∴ P = 1.04 × 10 −16 = 1.02 × 10–8 atm.


(ii) ∆G0 = –2.303 RT log Kp
∴ 29.29 ×103 = –2.303 × 8.31 × 300 log Kp
∴ Kp = 7.91 × 10–6

Illustration 13:
For the formation of ammonia the equilibrium constant data at 673K and 773K
respectively are 1.64 × 10–4 and 1.44 × 10–5 respectively. Calculate heat of the reaction.
Given R = 8.314 JK–1 mol–1.
Solution: Substituting the values in the equation
K p2 ∆H  T2 − T1 
ln =   , we get
K p1 R  T2 T1 

1.44 × 10−5 ∆H  773 − 673 


2.303log =
1.64 × 10 −4
8.314  773 × 673 
∆H(100)
2.303log(0.0878) =
8.314 × 773 × 673
100∆H = 2.303 (–1.0565) × 673 × 773 × 8.314
673 × 773 × 2.303 × −1.0565 × 8.314
whence, ∆H = = – 105216J = – 105.216kJ
100

You might also like