Lecture 6

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 59

Fundamental

equations of
Thermodynamics
&
Gibbs Free Energy

CHEH 240 Lecture 6


Fundamental equations of Thermodynamics

(1) The combined first and second law


From the first law: dU  dq  dW
dq
From the second law: dS 
T
dq
Where, dS  for irreversible system
T
dq
and, dS  for reversible system
T
For a closed system in which only reversible pV work
is involved dW   pdV and dS  dq
T
 dU  TdS  pdV Fundamental equation
The internal energy is a function of S and V
Where U, T, S, P, and V are state functions
2
 dU  TdS  pdV
The differential of U
 dU   dU 
 dU    dS    dV
 dS V  dV  S
Thus, we can calculate T and p as
 dU   dU 
T    and p   
 dS V  dV  S

S and V are natural variables of U represented as U(S,V)

3
The enthalpy was defined by: H  U  pV

by differential: dH  dU  pdV  Vdp


and dU  TdS  pdV
 dH  TdS  pdV  pdV  Vdp

 dH  TdS  Vdp
The natural variables of H are S and p represented
as H(S,p).
The last equation is the fundamental equation for H and for a
closed system in which only pV work, and since H is a state
function:

 dH   dH 
T    and V   
 dS p  dp  S

4
U and H provide criteria for whether a process can occur spontaneously in
a system when the corresponding natural variables are held constant.
dq
From: dS  , dW   pext dV
T
And substitute in : dU  dq  dW
We obtain:  dU  TdS  pext dV
At infinitesimal change (rev.) with constant S and V( dU )S ,V  0

“A change in a process can occur spontaneously if the internal energy


decreases when the change occurs at constant entropy and volume”
The meaning: dU = zero equilibrium
dU < zero spontaneous
dU > zero non-spontaneous
At constant S and p ( dH )S , p  0
dH = zero equilibrium
dH < zero spontaneous
dH > zero non-spontaneous 5
Gibbs and Helmholtz Free Energies
• System is in thermal equilibrium with surroundings at a constant
temperature. There is a change in the system accompanied by a
transfer of energy dq between system and surroundings. From the
Classius Inequality we have:
dq
ds  , i.e. Tds  dq or
T
dq  Tds  0
where the inequality applies when dq is for an irreversible process
and the equal sign applies when dq is for a reversible process.
This inequality may be developed in two ways:
1. Under the conditions of constant volume we have no
p-V work, i.e. in the absence of other forms of work:
dE  dq
 dE  Tds  0 6
2. Under the conditions of constant pressure, in the
absence of other forms of work other than p-V, we have :

dH  dq
 dH  Tds  0
The two inequalities above are so important (they determine
whether a process is going to occur or not) that we shall assign to
them two new thermodynamic functions namely:
A. The Helmholtz free energy, A, given by:
dA  dE  Tds
B. The Gibbs free energy, G, given by:
dG  dH Tds

7
Thus the two inequalities above governing the conditions
for a spontaneous process may be summarized as:

d A T ,V  0 from dE  Tds  0
d G T ,P  0 from dH  Tds  0

d A T ,V  0 & d G T ,P  0 when the system at equilibrium.

8
Helmholtz Energy (A)

It is defined by: A  U  TS

By differentiating: dA  dU  TdS  SdT


But  dU  TdS  pdV

 dA  SdT  pdV

T and V are the natural variables of A


If an infinitesimal change takes place in a system of constant
T and V, thus:
For irreversible process, A decrease.
For reversible process, A is constant.

It is more practical to use the criterion ( dA )t ,V  0


9
The Helmholtz energy can be used to show that the pressures of two
phases must be equal at equilibrium

For two-phase system in a container of fixed volume surrounded by a heat


reservoir
Suppose that the volume of phase α is increased by dV
and the volume of phase ß is decreased by dV. pß
So, the total volume is constant

 dA  SdT  pdV
When the system at equilibrium

dA = 0 = dAα + dAß

-pαdV +pß dV=0


p α= p ß

10
The Gibbs Energy (G)
It provides a more convenient thermodynamic
property than the entropy for applications of the
second law at constant T and p.
Example: for an isolated system consisting of system and
surrounding at constant T and p
Suniv  Ssys  Ssurr must increase for a spontaneous process
H sys at constant T
but S surr  
T
H sys
So that S univ   S sys
T
Multiply by -T  TSuniv  H sys  TSsys
-T∆Suniv = change in Gibbs free energy for the system = ∆Gsystem
For a process to occur spontaneously, ∆Gsys must decrease
This means that it is not required for specification what is
happening in the surrounding 11
Thus for a spontaneous process at constant T and p the Gibbs
energy must decrease. If the process is at equilibrium then dG = 0.
Thus for a spontaneous process at constant T and p the Gibbs
energy must decrease. If the process is at equilibrium then dG = 0.

This criteria is the most important developed so far because it leads to all
future analyses of spontaneous and equilibrium processes such as
equilibrium constants, electrode potentials and the Nernst equation, gas,
liquid and solid equilibria, solution processes, etc. etc.

While there are now two separate conditions for the two different paths
we will focus almost exclusively on the Gibbs Energy since the
constant pressure path is so much more convenient to establish
experimentally. We only need to do it in the open under constant
atmospheric pressure. Thus we will have tables of values for G rather
than A. Recall that this is exactly the same reason why we focus on H
and not U and why there are tables of H and not U.
Thus to determine if a constant T, p process will be spontaneous we
only need to find whether the change in one state function dG is
negative i.e. that the state function G decreases.
How does G behave overall as the state of a system changes?
Since the new criteria for spontaneous change is
that G must decrease, we can sketch how it must
behave as a system changes spontaneously. From
the plot we can see that the system naturally tends
to roll down a Gibbs energy hill until it reaches the
lowest point. The system is then at equilibrium. We
can see that a system tends to stay in a state G of
equilibrium because it must climb a Gibbs energy
hill to get out. Thus the reason for change, the
lowering of the Gibbs free energy can be viewed
as a “driving force” or a tendency for change.
The steeper the hill the greater the tendency for
the system to change.

The Gibbs (free) energy driving force is relatively simple in that it


is made up of two recognizable factors, the enthalpy and entropy. It is
useful in analyzing how each of these factors separately behaves
when a change occurs i.e. the Gibbs energy changes.
In general, any process in which the enthalpy (or energy)
decreases is favorable to a decrease in G and any process
in which the entropy goes up is also favorable to a
decrease in G.
In other words, systems, like most people, seek a
position of minimum energy and maximum disorder.
This leads to minimum Gibbs energy and a state from which a system
is reluctant to move.

Thus the Gibbs energy, enthalpy (energy) and entropy


are the three main properties to keep uppermost in
mind when thinking about a change of state.
Maximum non-expansion Useful Work
 The Gibbs free energy is the maximum amount of non-expansion
work that can be extracted from a closed system.
 We can show that for a constant temperature and pressure
process, a finite change in the Gibbs energy G is of great
physical significance because it is identical to the maximum work
that we can extract out of a system.
 This would be very useful in calculating the efficiency of a
process and would set an upper limit as to the maximum useful
work available.
 This has wide applications in assessing the efficiency of
processes and reactions.
Maximum non-expansion Useful Work
 H = U + pV, for a general change in conditions,
 the change in enthalpy is dH = dq + dw + d(pV)
 The corresponding change in Gibbs energy
( G = H – TS) is dG = dH – TdS – SdT = dq + dw + d(pV) – TdS – SdT
 When the change is isothermal, dT = 0; then dG = dq + dw + d(pV) – TdS
 When the change is reversible, dw = dwrev and dq = dqrev = TdS,
 For a reversible, isothermal process dG = TdS + dwrev + d(pV) – TdS = dwrev + d(pV)
 The work consists of expansion work, which for a reversible change is given by - p
dV, and possibly some other kind of work dwadd (e.g., the electrical work of
pushing electrons through a circuit or of raising a column of liquid).
 With d(pV) = pd V + Vdp, dG = (- pdV + d wadd,rev) + pdV + Vdp = dwadd,rev + Vd p
 If the change occurs at constant pressure (as well as constant temperature),
dp = 0 and, dG = d wadd,rev
 Therefore, at constant temperature and pressure, dwadd,rev = dG.
 However, because the process is reversible, the work done must now have its
maximum value, dwadd,max = dG 17
When a process is carried out in the open under constant
atmospheric pressure, pV work is done by the system against
atmospheric pressure [e.g. in expanding the volume of the system]
and this work is wasted because it is not controlled.

Thus any useful work done by the system is above and beyond this
pV work and is given by the change in the Gibbs energy.

For example in some electrochemical cells gases are evolved which


expand against the atmosphere to do pV work which is not used.
It is only the electrical energy generated, dwe,max equal to dG which is
utilized.
Thus the maximum useful net work that we can get from any
constant T and p process is

G = we,max = H - T S

18
ANALYSIS OF G IN TERMS OF H AND S
For useful work to be done, w, G and (almost always) H
If the entropy term TSmust be negative.
is positive [ then - TS will be negative] it
makes G more negative than just H and so increases the work
done for us.
In this case TS = qrev > 0 i.e. heat is transferred from the surrounding to
the system to help fuel the work.

In the reverse case where the system entropy decreases


TS < 0 [ -TS is positive and G is not as negative as H]
heat must flow from the system to the surroundings and so
all of the H is not available to do work..
The heat flowing to the surroundings increases the entropy of the
surroundings sufficiently so that the process remains spontaneous.

Thus part of the energy of the system has to be sacrificed in order to


maintain spontaneity.
19
20
Example: Check if this process is spontaneous or not?

NH4Cl(s) + H2O  aqueous solution at 25 oC

Hq(solution) = + 34.7 kJ mol-1 (endothermic) unfavorable to spontaneity

Sq(solution) = + 167.1 J K-1 mol-1 favorable to spontaneity

Gq(solution) =Hq (solution) - TSq (solution)


= 34.7 kJ mol-1 – 298 K (.1671 kJ K-1 mol-1)
= -15.1 kJ mol-1

In this case a favorable entropy change overcomes an unfavorable


energy change.

In fact, this process is spontaneous at any temperature


above:

21
22
Example: One mole of an ideal gas at 300 K and 10 atm is
isothermally and reversibly expanded to 1 atm. Calculate q,
w, U, H, S and G.

Analysis

23
24
STANDARD MOLAR GIBBS ENERGY
In the same way that standard heats of formation were defined for
compounds we define standard molar Gibbs free energy of formation.
This will allow us to calculate free energies of reactions at 298 K and
hence whether a reaction will be spontaneous IF the reaction is carried
out at constant
temperature and pressure.
fGq  Standard Gibbs (free energy) of formation of a compound [formed
from constituent elements in their standard state].
Values in tables are given for T = 298 K
fGq = 0 for elements in their standard state i.e. for O2(g), I2(s), C(s; graphite)
The standard free energy of the reaction is obtained in the same
manner as the heat of reaction and the entropy of reaction.

The standard Gibbs energy of formation of a compound and the


standard Gibbs energy of a reaction at constant temperature can also
be calculated from heats of reaction and entropies of reaction using:
Hq is almost independent of the temperature, Sq is
moderately dependent on the temperature
Gqis strongly dependent on the temperature 25
Example: oxidation of a-D glucose
C6H12O6(s) + 6 O2(g)  6 CO2(g) + 6 H2O(l)

If the Gq for CO2(g) = -394.4 kJ mol-1, Gq for H2O(g)= -237.2 kJ mol-1 , Gq
for C6H12O6(s) = -910.9 kJ mol-1. Is this reaction is spontaneous or not?

Calculate Gq = GqProduct - GqReactants


Gq =[ 6 Gq CO2(g) + 6 Gq H2O(l))]- [Gq C6H12O6(s) + 6 Gq O2(g)]
Gq = 6(-394.4) + 6(-237.2) - 1(-910.9) - 0 = -2879 kJ mol-1
If this reaction is carried out at constant temperature and pressure it would
be spontaneous.

Example: Iodine sublimes at 25 °C as I2(s)  I2(g) where the heat


and entropy of sublimation are H = 39.37 kJ mol-1 and S = 86.19 J
K-1 mol-1. What is equilibrium sublimation temperature if H and S
are assumed to be independent of temperature.

We use the fact that at equilibrium G = H – T S = 0

26
Given the following information, calculate the
standard Gibbs free energy of formation of H2O (g)
at 298.15 K.
Substance H o (kJmol1 ) S f (JK mol 1 )
o -1
f
H2O( g ) -241.818 188.825
H 2( g ) 0 130.684
O2( g ) 0
205.138

Solution: The standard Gibbs energy of formation is the


standard reaction Gibbs energy for the following reaction.
1
H 2 ( g )  O2 ( g )  H 2O( g )
2
G f ( H 2O( g ) )  H f ( H 2O( g ) )  TS f ( H 2O( g ) )
G f ( H 2O( g ) )  (241.818)  (298.15)(0.188825  0.130684  0.5 x0.205138)
kJ
G f ( H 2O( g ) )  228.572

mol 2
7
28
Maxwell Relations

(1) The combined first and second law


From the first law: dU  dq  dW
dq
From the second law: dS 
T
dq
Where, dS  for irreversible system
T
dq
and, dS  for reversible system
T
For a closed system in which only reversible pV work
is involved dW   pdV and dS  dq
T
 dU  TdS  pdV Fundamental equation
The internal energy is a function of S and V
Where U, T, S, P, and V are state functions
29
 dU  TdS  pdV
The differential of U
 dU   dU 
 dU    dS    dV
 dS V  dV  S
Thus, we can calculate T and p as
 dU   dU 
T    and p  Equation I
 dS V  dV S

S and V are natural variables of U represented as U(S,V)


Also since U is a state function we can apply the state function condition

   U      U    T   p 
             
 V  S V S 
 
S V S  V  V S  S  V
This is a Maxwell Relation
30
Using the fundamental equation we can obtain new
equations for dH, dA and dG and use the same
procedures as above to obtain three more equations like
[I] as well as three other Maxwell Relations. These turn
out to be particularly useful in manipulating partials as
will be shown shortly.
H = U + pV and G = H – TS
For a constant T and p process dT = 0 and dp = 0 and thus
G = U – TS + pV dG = dU – TdS –SdT + pdV + Vdp

Substitute the fundamental equation dU = TdS – pdV into the equation for
dG to get: dG = Vdp – SdT
This equation for dG suggests that we take p and T as the variables
for the Gibbs energy.

 G   G 
dG    dp    dT
Comparing coefficients of dp and
 p T  T p dT for the two equations gives

 G   G 
   V and    S Applying the state function condition for G
 p  T  T  p
 V   S 
      This is a Maxwell Relation
 T p  p T
The four Maxwell relations are

 p   S   V   S 
        
 T V  V T  T p  p T
 T   V   T   p 
        
 p S  S P  V S  S V

Maxwell relations allow us to develop different equations


Start with the fundamental equation dU = TdS – pdV.
Proof that  U   T  p   p
   

 T
V  T  V
To get the LHS partial of the above equation we divide dU by dV
and hold T constant
dU  U   S 
   T  p
dV T  V T  V T
 p   S 
Substitute the Maxwell relation  T     to obtain the equation
  V  V T

 U   p 
   T   p
 V T  T  V

The physical significance of the LHS. It is the change in the energy when
the volume of a system, say a gas, is changed. i.e, when the distance
between gas molecules is increased or decreased. Thus it is a measure
of the change in the potential energy of molecules.
THE CHANGE of GIBBS ENERGY WITH T AND p

It is easy to determine the spontaneity of a reaction at 1 atm and 298 K


since we can get the free energy of a reaction from the free energy of
formation of compounds given in tables.

But, there are many reactions that are carried out at very different
temperature and pressure conditions.
It is important to know how the free energy changes
with temperature and pressure if we want to
determine spontaneity at any temperature and
pressure.

This is especially true for temperature changes since G is strongly


dependent on temperature for solids, liquids, gases and solutions.

Pressure effects on G are substantial only for gases.


35
36
Change of G with temperature
 G 
The change of G with T was given by    S
 T  p

This is the slope of the plots of G vs T

The fact that the slope is just


the negative of the entropy
makes the interpretation of
the plots very simple.

The entropy of a gas is much larger than that of a liquid


which in turn is larger than that of the solid. Thus, it is
easy to rationalize the relative steepness of the G vs T
lines for gas, liquid and solid.
37
 G 
For the change of G with T we can write    S
 T  p
“what will happen to the equilibrium which exits between
water  steam at 100 °C and 1 atm if we raise the
temperature to 110 °C”?
It is known that: at 100 °C, 1 atm there is equilibrium and
Gq = Gq(steam) - Gq(water) = 0.
We need to see whether the value of Gq becomes positive or
negative when the temperature increases by +10 °C.
 G 
From    S , write the equation for a small finite change d as
 T p
q
 G  dG q q q
    S  dG  (S )dT
 T  p dT
Since Sq [ Sq(steam)> Sq(water)]  0 and that dT  0, dGq < 0 at 110
°C, 1 atm and hence , the process water  steam will be spontaneous.
38
a) The Gibbs Helmholtz Equation
dG q
 Sq  dG q  (Sq )dT
dT
The above temperature dependence of the Gibbs energy on entropy is OK but it would be
even better if it depended on enthalpy since we usually have more heat data than entropy
data. This can be shown as follows:
HG  G 
G  H  TS  S    
T  T p
 G/T 1  G    1/ T  1  G  G
Consider       G      
 T p T  T p  T p T  T p T 2

 G  GH  G/T H
Substitute for  T     T   T 2
 p T  p

For any process: initial state  final state


with G = Gf – Gi and H = Hf - Hi
  G / T   H
   This is Gibbs Helmholtz Equation
 T  p T2 39
 G 
d T  H G dT
     d    H 2
 T P
2
 dT  T T
 P
 dT 1  1 1 dT
    d     2
 dT P T2 T  T
G 1 GT2 G T2 1
 d    Hd  
 T P T 
GT1
d    H  d  
T  T1
T 
GT2 GT1 1 1
  H   
T2 T1  T2 T1 
GT2 GT1 1 1 GT2 G298 1 1 
  H      H   
T2 T1  T2 T1  T2 298  2
T 298 
This equation is useful because if we know ∆G for a reaction as a
function of temperature, the enthalpy change for the reaction can
be calculated or if ∆H and ∆G are known at one temperature, ∆G
can be calculate at another temperature assuming that ∆H is
independent of temperature
29
The standard enthalpy of the reaction N2 + 3H2  2NH3 is -92.2 kJ
mol-1 & the standard Gibbs function is -31.0 kJ mol-1, both at 298
K. Estimate the Gibbs function at (a) 500K & (b) 1000K. Is the
reaction spontaneous at room temperature? Is the formation of
ammonia favored or disfavored by a rise in temperature?
GT2 G 298

1 1 
  H   
T2 298  T2 298 
G500   31.0kJ .mol 1 
    1
   92.2kJ .mol 1  
1 
500 K  298 .15 K   500 K 298 K 
G500  10.4kJ .mol 1
G1000   31.0kJ .mol 1 
  
   92.2kJ .mol 1 1000
1

1 
K 298 K 
1000 K  298 .15 K   
G1000  113 .2kJ .mol 1

Comments: With increasing temperature the free energy


Increases, indicating that the process is not favored with
increasing temperature.
30
Change of G with pressure

 G 
It was proved before that    V
 p  T
The slope of the plot of G vs p is just the
volume of the system. Since V is always
positive, the free energy must always
increase with pressure.
The slopes of the plot for the gas, liquid and solid should decrease in that
order since the molar volumes of these phases decreases as we go from
gas to liquid to solid.
 G 
For a finite change in G when p changes :    V
 p  T
This equation can be applied to the water  steam equilibrium at 100
°C, 1 atm.
What will happen to the equilibrium if the pressure is decreased to
0.5 atm?
 G  dG
    V dG  (V) dp
 p T dp
42
BUT V = V(steam) – V(water) >> 0 and dp = 0.5 – 1.0 = - 0.5 < 0
and as a result dG < 0

Water turning into steam (water vapor) will occur spontaneously


Deriving equation for the change of Gibbs energy with pressures at
constant temperature:

From the above partial derivative, at constant temperature:


 G  dG
    V and dG |T  Vdp |T
 p T dp T
Thus  dG   Vdp the understanding that for this integration, T is constant

pf

G(pf )  G(pi )   V(p,T)dp Always true for any isothermal process


pi

43
pf

G(pf )  G(pi )   V(p,T)dp


pi

a) For solids and liquids ( V is independent of p)

G 2  G1  V(p2  p1 ) G  G  V(p  p )
o o

Where, Go and po are the standard value


Since for most solids and liquid, the molar
volume is small the value of G is also very
small unless the pressure change is huge.

Thus as a first approximation we can say that


G is independent of pressure for solids and
liquids. This is seen in the plot of G vs p
where the lines for liquid and solid are almost
flat.

44
pf

G(pf )  G(pi )   V(p,T)dp


pi

b) For gasses ( V is dependent of p), ideal gas


G2 p2
nRT dp
V  G  nRT  d
P G1 p 1

p2
 G 2  G1  nRT ln
p1
V1 V2
 G  nRT ln  nRT ln  W
V2 V1
p
G  G  nRT ln o
o

p
This equation shows that at constant temperature the free energy goes up with
the pressure. This means that for an isothermal expansion the Gibbs energy
decreases (due to a dispersal of energy)
45
The effect on G of a gas due to changing the pressure
is much greater the the effect on G of the
corresponding liquid or solid, because the molar
volume of the gas is much larger.

Example: Calculate the free energy change G when one mol of


an ideal gas at a constant temperature of 300 K is compressed
from 1 atm to 100 atm.

Analysis:
G should be positive since this is a compression which concentrates energy.

G = nRT ln(pf/pi)
= 1 mol x 8.314 J K-1 mol-1 x 300 K ln(100/1) = 11.5 kJ

46
Example: fGO (Standard Gibbs energy of formation) for liquid CH3OH
at 298K is -166.27 kJmol-1, and that for gaseous CH3OH is -161.96
kJmol-1 . The density of the liquid methanol at 298K is 0.7914 g.cm-3.
Calculate fG (CH3OH, g) and fG (CH3OH, liq ) at 10 bar at 298 K and

fG (CH3OH, g)
fG = fGo +nRT ln(p/po)
= -161.96 kJmol-1+(1 mol x 8.314J K-1 mol-1x 298K ln(10/1)
= -156.25 kJmol-1

fG (CH3OH, liq) 1 bar = 100 kPa


fG = fGo + Vm(p-po) and

Vm= molar mass/ density = 32 g mol-1 / 0.7914 g.cm-3= 40.49 cm3 mol-1
Vm= 40.49 x 10-6 m3 mol-1
Vm(p-po) = 40.49 x 10-6 m3 mol-1 (10 x 105- 1 x 105)Pa x (1 x 10-3 kJJ-1)
fG = -166.27 kJmol-1+ 40.49 x 10-6 m3 mol-1 (10 x 105- 1 x 105)Pa x (1 x
10-3 kJJ-1) = -166.23 kJmol-1

47
Example: An ideal gas at 27oC expands isothermally and reversibly
from 10 to 1 bar against a pressure that is gradually reduced. Calculate
q, W, U, H, G, A, and S.

Analysis: Isothermally and reversibly, U, H, equal zero


Gas expand, W is negative, q = -W, A is negative, S is positive
and G is negative
Calculations:
Wmax = A = -RTln(p1/p2) = -(8.314 JK-1 mol-1)(300 K) ln (10/1)
= -5746 J mol-1
q = - Wmax = 5746 J mol-1
U = H = 0
S = (qrev/T )= (5746 J mol-1 )/(300 K) = 19.14 J K-1 mol-1
G = H - T S= 0 – (300 K) (19.14 J K-1 mol-1 ) = 5746
Jmol-1
OR G = RTln(p2/p1) = (8.314 JK-1 mol-1)(300 K) ln (1/10) = -5746
Jmol-1
48
INTRODUCTION OF THE CHEMICAL POTENTIAL

The change in number of moles of any substance must be considered.


Starting with pure substances and determine how the Gibbs function
will change if infinitesimal amount of the same substance is added at
constant temperature and pressure.  G 
This partial derivative is defined as the Chemical Potential :    
 n  T , p
For a pure substance ,  is just the molar free energy Gm.

 nG m   G m 
G = n Gm    Gm  n  
 n  T , p  n  T , p
But for a pure substance Gm is constant with respect to the n and thus 
= Gm.
This is true of any molar quantity.

For example, the molar volume Vm of water is 0.018 L mol-1 and is


independent of whether we are talking about 1 mole or 23 moles. Molar
quantities are “intensive” like density - they don’t depend on the amount.
49
The chemical potential is the most fundamentally important
function for general systems because it indicates how a system
will change with a change in temperature, pressure AND
composition.

As applied to ideal gases we can obtain the chemical potential of an


ideal gas at any temperature and pressure p,T from the previous
equation for G:

q p q p
    RT ln q
OR     RT ln
p 1 atm
In the most general formulation  is a function of T, p and moles of
each component in the system ie
 =  (T, p, n1, n2, n3, ………)

Also rewrite the equilibrium criteria for a constant T and p process


G = 0 as  = 0
50
FUGACITY

It is simply a measure of molar Gibbs energy of a real gas .


Modify the simple equation for the chemical potential of an ideal gas
by introducing the concept of a fugacity f. The fugacity is an “
effective pressure” which forces the equation below to be true for real
gases:
q  f 
(p, T)   (T)  RT ln q  where pq = 1 atm
p 
A plot of the chemical potential for an ideal
and real gas is shown as a function of the
pressure at constant temperature.

The fugacity has the units of


pressure. As the pressure
approaches zero, the real gas
approach the ideal gas behavior
and f approaches the pressure.
51
If fugacity is an “effective pressure” ie the pressure that gives the right value
for the chemical potential of a real gas. The only way we can get a value for
fugacity and hence for  is from the gas pressure. Thus we must find the
relation between the effective pressure f and the measured pressure p.

let f=fp
f is defined as the fugacity coefficient. f is the “fudge factor” that
modifies the actual measured pressure to give the true chemical potential
of the real gas. q  f 
(p, T)   (T)  RT ln q 
p 
By introducing f we have just put off finding f directly. Thus,
now we have to find f. Substituting for f in the above equation
gives:  p 
q
(p, T)   (T)  RT ln  q   RT ln f  (ideal gas)  RT ln f
p 
(p, T)  (ideal gas)  RT ln f
This equation shows that the difference in chemical potential between
the real and ideal gas lies in the term RT ln f. This is the term due to
molecular interaction effects.
52
The equation relating f or f to the measured pressure p:

Note that as p  0, the real gas  ideal gas, so that fp and f 1
The chemical potential for an ideal gas and a real gas at two pressures p
and p is
p
p

p Vp ideal,m dp   dideal  ideal (p, T)  ideal (p , T)  RT ln  p 



f 
 V dp   d  (p, T)  (p, T)  RT
p
m ln  
 f
Subtracting the first equation from the second gives
p
p
f 
 V
p
m  Videal,m  dp  RT ln   RT ln  
 f  p 

 f /p 
p

or ln 
   
1
  Vm  Videal,m  dp
 f / p  RT p
Let p  0, then in the initial state the real gas  the ideal gas. Thus f  p

f 
p

ln   
1
  Vm  Videal,m  dp
 p  RT 0
53
RT  RT 
Since for an ideal gas Videal,m  and for a real gas Vm   Z
p  p 

where Z is the compressibility factor

f  1  RT   RT    Z(p, T)  1
p p

ln      
 p  RT 0  p 
Z   
 p 
dp  
0
 p 

dp  ln f

 p  Z(p, T)  1  
f  p exp     dp 
 0  p  

The fugacity coefficient f = f/p is given by

 p  Z( p , T )  1  
f  exp    dp 
0  p  
Thus the fugacity of a gas is readily calculated at same pressure p if Z is
known as a function of pressure up to that particular pressure.
At low to intermediate pressures Z<1, attraction dominates and
the area under the curve of the RHS plot, Z  1 dp
negative.
 p is

Thus f =p exp(negative power). and f < p, f<1.

Z 1
At high pressures Z>1, repulsion dominates and  dp >0 and f>p, f>1
p

55
The Fugacity of van der Waals

From the compressibility factor Z of a Van der waals:

 pVm    a  p 
Z   p
RT a
 2
 RT  Z  1  b    
Vm  b Vm   RT    RT 
p
 Z 1

From the fugacity equation f
ln   dp
0
p p 
p
  a  1   a  p

f
ln   b     dp   b    
p
0   RT   RT   RT   RT
  a  p 
f  p exp b     
  RT   RT 
56
Example: Given that the van der Walals constants of nitrogen
are: a= 1.408 L2 atm mol-2 and b=0.03913 L mol-1.
Calculate the fugacity of nitrogen gas at 50 bar and 298 K.

f   a  p

ln   b    
p   RT   RT

  a  p
ln f  ln p   b    
  RT   RT

  1.408L2 .bar .mol 2  50bar


 1 
ln f  ln 50  0.03913L.mol   
 1 1  1 1
  (0.08314L.bar . K .mol )( 298K )   (0.08314L.bar . K .mol )( 298K )

fugacity (f) = 48.2 bar

57
Fundamental equations for open system
An open system is the one that can exchange matter with its surroundings.
Consider a two component system at temperature T and pressure p
which is composed of n1 moles of substance 1 and n2 moles of
substance 2.
The Gibbs function can be written as G=G(T,p, n1, n2). It is a function of four
independent variables.
 G   G   G   G 
The total differential is dG    dT    dp    dn    dn 2
   
1
 T p,n1 ,n 2  p T,n1 ,n 2  1 p,T,n 2
n  2 p,T,n1
n

For a system of constant composition ie a closed system:


G  G 
dG = -S dT + V dp and    S   V
 T p,n ,n 1 2 p T,n ,n 1 2

 G   G 
Also from the earlier definition of the chemical potential
 :   1    2
 n1  p,T,n 2 
 2 p,T,n1
n

Gibbs function for a two component open system can be


written as
dG = -S dT + V dp + 1dn1 + 2dn2 = -S dT + V dp +idni
58
This is the most general expression for the driving force of a
system. It is very useful in examining what happens in a system at
constant temperature and pressure when the amounts of n1 and n2
are changed. This occurs frequently in chemical and biochemical
processes for example in the transfer of a solute to a solution. The
starting equation in this case where dT = 0 = dp is
dG = 1dn1 + 2dn2 = dwe,max
This equation is also interesting because we know that the reversible free
energy change at constant temperature and pressure is the maximim
non pV work. This equation shows that non pV work is the result of
composition changes in the system.

This is exactly what happens in a electrochemical cell. As the


electrochemical reaction occurs there are compositional changes and
these lead to electrical current.
Finally, if a process is at equilibrium then dG = 0 and
1dn1 + 2dn2 = 0 or 1dn1 = - 2dn2
This equation known as the Gibbs-Duhem equation is the starting point
for many developments where equilibrium compositional changes occur
under constant temperature and pressure. 59

You might also like