Vous êtes sur la page 1sur 24

121

Helmholtz and Gibbs Free Energies



We have discussed at some length the Second Law of Thermodynamics
which provides criteria for spontaneity and reversibility applicable to any
processes.
S
Total
= S + S
Surr
> 0 for a spontaneous process and S
Total
= 0 for a
reversible process or for conditions of equilibrium. Such criteria although
very useful are not extremely practical as they require us to carry out
calculations for both the system and the surroundings. It would be
extremely advantageous to find criteria for spontaneity which are based on
the use of thermodynamic state functions for the system and which do not
invoke the surroundings. This can be done under specific conditions
(constant V and T or constant P and T) which are encountered in many
processes in chemical, physical and materials sciences.

Helmholtz Free Energy:
Let us assume we are dealing with a process occurring at constant V. The
heat exchanged with the surroundings is equal to change in internal energy.
dU = q
V

122
The second law, written in the form of the Clausius inequality states:
dS > q / T
Note that this inequality comes from dS + dS
Surr
> 0 and dS
Surr
= - q / T
Therefore, we can rearrange this equation as q - T dS < 0 since T > 0.
This leads in the case of a process at constant V to dU - TdS < 0
If we now assume T to be constant (i.e. dT = 0), we get: d(U - TS) < 0
We then define a new thermodynamic function, A, the Helmholtz free
energy such that:
A = U - TS
The criterion for spontaneity for a process occurring at constant V and T is
therefore dA < 0. In contrast, a process occurring at constant V and T will be
reversible or will have reached some equilibrium state if dA = 0.

Gibbs Free Energy:
We now consider processes occurring at constant P. For constant P
processes, the heat exchanged with the surroundings is equal to the change
in enthalpy.
q
P
= dH
Therefore the Clausius inequality for a spontaneous process taking place at
constant P is written as:
123
dS > dH / T which is equivalent to:
dH - T dS < 0
If the process furthermore occurs at constant T then:
d(H-TS) < 0
We then define a new thermodynamic state function, G = H - TS, which we
call the Gibbs Free Energy.
The criterion for spontaneity for a process occurring at constant P and T is
therefore dG < 0. In contrast, a process occurring at constant P and T will be
reversible or will have reached some equilibrium state if dG = 0.

Maximum Work Functions, A and G
Besides providing criteria for the spontaneity of processes occurring either at
constant V and T or P and T, the Helmholtz and Gibbs Free Energies have
further significant meanings.
Let us consider first the case of the Helmholtz free energy.
The second law states that dS q / T. Using the first law dU = w + q, we
get:
T dS dU - w which is equivalent to: w dU - T dS.
If the process occurs at constant temperature, then we can write:
w dA or w A
124
If we focus on a process whereby work is being done by the system
(w < 0). We now recall from the discussion of reversible vs. spontaneous
expansions of ideal gases that the maximum work of expansion is done by
the system when the expansion occurs reversibly. This corresponds to the
equality between w and A. We therefore conclude that:
A is the maximum work that can be done by a system at constant T.
This is actually why the symbol A was chosen for the Helmholtz Free
Energy. In german, A stands for Arbeit, which in english means work.
Let us consider now the meaning of G.
Using a similar approach, we write the first law in its most general form.
dU = w + w
e
+ q
where w is the volume expansion work and w
e
is any extra (non-
expansion) work.
We now use the second law in the Clausius formulation.
dS q /T
which leads to:
T dS dU - w - w
e

which for a constant pressure process (P = P
ext
= cst and dP = 0), we can
rearrange as:
125
w
e
dU + P dV - T dS = dU + P dV + V dP - T dS = dH - T dS
Therefore for a constant T (dT = 0) and P process:
w
e
dH - T dS - S dT = dG
w
e
G
The maximum work (i.e. work done under reversible conditions) by a system
when P and T are constant is the change in free energy of the process.
This statement finds direct applications in a number of situations, such as in
the use of batteries (which are system based on electrochemical reactions
producing current (i.e. electrical work)).

Standard Free Energies of Reactions
In the same way as we defined standard reaction enthalpies, we can define
standard reaction entropies and standard reaction Gibbs free energies. We
focus on standard reaction Gibbs free energies rather than on standard
reaction Helmholtz free energies because chemical reactions are easier to
carry out at constant pressure and temperature, than at constant volume and
temperature.
The standard reaction entropy, defined as:
R
S

(T) is given by:

R
S

T
=
J

J
S
m

T
(J)
126
where
J
is the stoichiometric coefficient in front of substance J in the
reaction equation and S
m

T
(J) is the standard molar entropy of substance J
at the temperature T (and at the standard pressure P

)
Similarly, we define the standard reaction Gibbs Free Energy by:

R
G

T
=
J

F
G

T
(J)
where
F
G

T
(J) is the standard formation Gibbs Free Energy for the
substance J at temperature, T and standard pressure, P

. Note that the


standard reaction free energy of an element in its reference state is equal to
zero (the same as was discussed for standard reaction enthalpies).
We obviously also have the relationship:
R
G

T
=
R
H

T
- T
R
S

T








Chapter V: The Second Law: The Machinery

127
In the previous chapter, we introduced two new thermodynamic state
functions, A and G. These functions were defined by:
A = U - TS and G = H - TS and obviously as for H and U, these functions
are extensive and have units of energy (i.e. Joule). For these functions, we
can also define the corresponding molar quantities, A
m
and G
m
by:
A
m
= A / n and G
m
= G / n.
Note however, that these definitions for the molar free energies are only
valid for pure, one component systems. More general definitions will be
given later when we deal with mixtures.

For a given process, it will be important to know how to calculate either the
change in the Gibbs free energy or the change in the Helmholtz free energy,
because as was thoroughly discussed earlier, these quantities tell us whether
a given process is spontaneous or has reached equilibrium. Furthermore,
these quantities enable us to determine the maximum work that a given
system can do under specific conditions.


Change in the Internal Energy: dU
We first write the first law as:
128
dU = w + q
Now, we recall that the change in energy (a state function) is independent of
the path chosen, therefore the change in U can be calculated using a
reversible path, assuming, of course, we know what the initial and final
states are (so we can define that reversible path). For a reversible path, we
can write: w = - P
ext
dV = - P dV
The second law for a reversible process states that dS = q / T. We can
therefore write dU:
dU = - P dV + T dS
This very important equation is known as the fundamental equation of
thermodynamics.
U is a state function, thus, dU is an exact differential and we can write:
U
V






S
= P and
U
S






V
= T
For an exact differential, the mixed second derivatives are independent of
the order of differentiation, thus we can write:
129

S
U
V






S






V
=

V
U
S






V






S
Therefore :
P [ ]
S






V
=
T [ ]
V






S

P
S






V
=
T
V






S

This last equation is known as the First Maxwell Relation. It provides a new
relationship between materials parameters. Indeed, taking the reciprocal of
this equality tells you that the change in volume with temperature during an
adiabatic reversible process (change in V with T at constant S) is equal to
minus the change in disorder accompanied by a change in pressure during an
isochoric experiment (change in S with P at constant V).

Change in Enthalpy: dH
We know H = U + PV, therefore dH = dU + P dV + V dP
We use the expression of dU derived above (dU = - P dV + T dS)
and obtain:
dH = - P dV + T dS + P dV + V dP and get:
dH = V dP + T dS
130
Again, since H is a state function, dH is an exact differential. The terms V
and T correspond to the first derivatives of H with respect to P and S at
constant S and P, respectively.
H
P






S
= V and
H
S






P
= T
Since for an exact differential, the mixed second derivatives are independent
of the order of differentiation, we can write:

S
H
P






S








P
=

P
H
S






P








S
Therefore :
V [ ]
S






P
=
T [ ]
P






S
V
S






P
=
T
P






S

This last equation is known as the second Maxwell Relation. It provides a
new relationship between materials parameters. Indeed, it tells you that the
change in pressure with temperature during an adiabatic reversible process is
equal to the change in disorder (entropy) resulting from a change in volume
during an isobaric experiment.



131
Change in Helmholtz Free Energy: dA
We know that A = U - TS, therefore dA = dU - T dS - S dT
We use the expression of dU derived above and obtain:
dA = - P dV + T dS - T dS - S dT and get:
dA = - P dV - S dT
Again, since A is a state function, dA is an exact differential. The terms -P
and -S correspond to the first derivatives of A with respect to V and T at
constant T and V, respectively.
A
V






T
= P and
A
T






V
= S
Since for an exact differential, the mixed second derivatives are independent
of the order of differentiation, we can write:

T
A
V






T








V
=

V
A
T






V








T
Therefore :
P [ ]
T






V
=
S [ ]
V






T
P
T






V
=
S
V






T

This last equation is known as the third Maxwell Relation. It provides
another relationship between materials parameters. Indeed, it tells you that
the change in pressure with temperature during an isochoric process is equal
132
to the change in disorder (entropy) accompanied by a change in volume
during an isothermal experiment. Note that the term on the left hand-side is
the ratio of the isobaric coefficient of thermal expansion () to the
isothermal compressibility coefficient (
T
).

Change in Gibbs Free Energy: dG
We know G = H - TS, therefore dG = dH - T dS - S dT
We use the expression of dH derived above and obtain:
dG = V dP + T dS - T dS - S dT and get:
dG = V dP - S dT
Again, since G is a state function, dG is an exact differential. The terms V
and -S correspond to the first derivatives of G with respect to P and T at
constant T and P, respectively.
G
P






T
= V and
G
T






P
= S
Since for an exact differential, the mixed second derivatives are independent
of the order of differentiation, we can write:
133

T
G
P






T








P
=

P
G
T






P








T
Therefore :
V
[ ]
T






P
=
S
[ ]
P






T
V
T






P
= -
S
P






T

This last equation is known as the fourth Maxwell Relation. It provides an
additional relationship between materials parameters. Indeed, it tells you that
the change in volume with temperature during an isobaric process is equal to
minus the change in disorder (entropy) accompanied by a change in pressure
during an isothermal experiment. Note that the term on the left hand side is
related to the isobaric coefficient of thermal expansion.

The above considerations led to the development of four new and important
relationships. It is however, possible to recall each of these relationships
quickly without much derivation. Indeed, note that in each case the
thermodynamics variables or functions (P, V, T, S) work in pairs. P is
always associated with V and S is always associated with T. Indeed this is
expected as the work involves P and V and the heat involves S and T. The
quantities P and V, on one hand, and S and T, on the other, are called
conjugate pairs. Also, note that the Maxwell relations involve partial
134
differentials of the type
M
Y






N
and
X
N






Y
where M and N form a
pair of conjugate variables and X and Y are the other pair of conjugate
variables. Note that Y and N, on one hand and X and M on the other hand
are not pairs of conjugate variables.
A method to quickly obtain the Maxwell relations is illustrated below:
Let us assume we wish to know the Maxwell equation involving
P
S






V

First, we know that the only variables involved in a Maxwell equation are P,
V, S and T. Second, we know that P is conjugate with V and S is conjugate
with T, so we place V and T diagonally from P and T.
P
S






V
=
T
V

Then we introduce the partial differential symbols:
P
S






V
=
T
V







Then we recall that if the first partial corresponds to a derivative with respect
to S at constant V (in the L.H.S. term) then in the R.H.S. term, we must have
a derivative with respect to V at constant S. Therefore we introduce S as a
subscript for the second partial derivative, to indicate that it is taken at
constant S. We also check for the signs of each of these partials to make
135
sure the relationship is physically meaningful. Remember V increases when
T increases, P increases when T increases, S increases when V increases and
S decreases when P increases. In the specific example chosen here, we must
introduce a negative sign. So we finally get:

P
S






V
=
T
V






S

In some case, you may need to estimate a partial differential, which at first
sight, looks like it may be involved in a Maxwell relation (because it
involves P, V, S and T and because P and S do not form a conjugate pair),
but is not exactly one. For example consider,
P
S






T

It is not exactly involved in a Maxwell relationship because P and T do not
form a conjugate pair. In this case, all you have to do is to examine the
reciprocal of the desired partial derivative (i.e.
S
P






T
), which is involved in
a Maxwell relation. You obtain:
S
P






T
=
V
T






P

So you conclude that:
P
S






T
=
T
V






P

136
On the other hand, the quantity
P
V






T
can never be related to a Maxwell
relationship, because P and V form a conjugate pair. In this case, this partial
differential is already related to a materials property it is equal to
(-1/
T
V). The same holds for
S
T






P
which is equal to C
P
/ T. You may
want to convince yourself that you can prove this!!

The Internal Pressure and (H/P)
T

When we discussed the change in internal energy with volume and
temperature and the change in enthalpy with pressure and temperature, we
defined dU and dH by the following expressions:
P T
T
V T
T
T P
T V
T
V
T V
P
H
and P
T
P
T
V
U
where
dP dT C dH
dV dT C dU

=
+ =
+ =



We are now in a position to derive these two relations. For the internal
pressure, which is the partial differential of U with respect to V at constant
T, we write:
dU = PdV + TdS
dU
dV
=
PdV
dV
+
TdS
dV

dU
dV
= P + T
dS
dV

137
We transform the ratio of exact differentials in ratio of partial differentials
and consider those to be obtained at constant T (as in the expression of the
internal pressure) and get:
U
V






T
= P + T
S
V






T

We now make use of the corresponding Maxwell relation:
S
V






T
=
P
T






V

and obtain the desired relationship:
T
=
U
V






T
= T
P
T






V
P
Similarly, for the expression of the partial of H with respect to P at constant
T, we start with dH
dH = VdP + TdS
dH
dP
=
VdP
dP
+
TdS
dP

dH
dP
= V + T
dS
dP

We transform the ratio of exact differentials in ratio of partial differentials
and consider those to be obtained at constant T and get:
H
P






T
= V + T
S
P






T

We now make use of the corresponding Maxwell relation:
S
P






T
=
V
T






P

138
and obtain the desired relationship:
H
P






T
= V T
V
T






P


Temperature Dependence of the Gibbs Free Energy:
Gibbs-Helmholtz Equation
We note that G is a function of P and T. Any process can be described as a
combination of two processes where, in the first one, the pressure is changed
at constant temperature (isothermal) and in the second, the temperature is
changed at constant pressure (isobaric). It is therefore very valuable to know
how G changes with P (at constant T) and with T (at constant P).
We will first examine the temperature dependence of the Gibbs Free Energy
at constant pressure. If P is constant, dP = 0. Therefore:
dG = - S dT
Since the entropy is always positive and increases with temperature, the free
energy of any pure substance decreases with increasing temperature at a
faster rate as the temperature is higher (i.e. G is a decreasing function of
temperature with a concave down shape).
Diagrams of G vs. T will be investigated further as we discuss phase
transitions (next chapter).
139
To discuss the temperature dependence of G more quantitatively, we use the
definition of entropy from the expression of dG.
S =
G
T






P

Therefore we can rewrite G = H - TS as:
G = H + T
G
T






P
H = G T
G
T






P

Dividing on both sides by - T
2
we get:
H
T
2
=
G
T
2
+
1
T
G
T






P
= G

1
T






T












P
+
1
T
G
T






P
=

G
T






T












P

The final result is shown below and allows us through integration to estimate
the variation of G with temperature provided we know the enthalpy of the
system.
( )
( )
( )
( ) ( )
constant considered be can H if
1 1
1 2 1
1
2
2
2 2
2
1
2
2
1
1


T T
H
T
T G
T
T G
dT
T
T H
T
G
d
T
H
T
T
G
T
T
T
T G
T
T G
P




140
Pressure Dependence of the Gibbs Free Energy at Constant T:
(1) Cases of solids or liquids.
In most cases we can assume condensed phases (liquid and solids) to be
incompressible (assume that their volume is independent of pressure).
Obviously, this is a good approximation if you are not dealing with
processes where P can reach extremely high values (geological sciences).
dG = V dP (since T is constant, dT = 0)
Then: G = V P
The free energy of a liquid increases almost linearly with pressure (almost,
because, rigorously, V decreases very, very slightly when P increases).

(2) Cases of Ideal Gases: V = nRT / P
In this case, we can integrate dG = nRT (dP/P) as:
G = nRT ln (P
F
/ P
I
) since T is constant.
Note that this result can be obtained another way starting with G = H - TS
dG = d(H - TS) = dH - T dS (since T is constant).
Therefore, G = H - T S
We showed earlier that for the isothermal expansion or compression of an
ideal gas (change in P at constant T) that dU = 0 and dH = 0, therefore
H = 0
141
dS = q / T = - w / T = +P dV / T => S = nR ln (V
F
/ V
I
) Thus,
S = -nR ln (P
F
/ P
I
)
Therefore, G = nRT ln (P
F
/ P
I
)
We will often come back to the above equation to describe the molar free
energy of an ideal gas. Specifically, we will make great use of the
relationship between the free energy of one mole of an ideal gas at some
temperature T and pressure P and one mole of the same gas at the same
temperature but at the standard pressure, P

. We can then write:


G
m
= G
m
(T, P) - G
m
(T, P

) = RT ln (P / P

) or
G
m
(T, P) = G
m
(T, P

) + RT ln (P / P

)
Because the molar free energy of a substance is a central quantity in many
thermodynamic applications (mixing, chemical reactions, etc...), we will
give it another name: the chemical potential and use the symbol to denote
it. We define the chemical potential of a substance in the most general way
(accounting for cases where the system contains many different substances)
through the following relation:

J
=
G
n
J






P,T,n
I
n
J

The chemical potential of a substance is therefore the change in free energy
of a system when you add one mole of substance J to that system, keeping
142
the temperature, the pressure and the number of moles of any other
substances present in the system the same. We will come back to this topic
when talking about mixtures and chemical reactions....
The chemical potential of an ideal gas is therefore generally written as:
T, P ( ) =

T, P

( )
+ RT ln
P
P









(3) Cases of Real Gases:
Since the gas is not ideal, we cannot use V = nRT / P. However, we really
like the equation describing the chemical potential (i.e. the molar free
energy) of an ideal gas. So, we will use the same kind of equation, but
modify it a little, because we know that the above equation is not valid for
real gases. We will define a new quantity, f, called the fugacity, which has
no physical meaning but which is very practical. We will also define a
dimensionless quantity, called the fugacity coefficient, = f / P. The
fugacity has units of pressure.
T, P ( ) =

T, P

( )
+ RT ln
f
P








To be able to calculate the fugacity, we will contrast the change in chemical
potential for both a real gas and an ideal gas as we change the pressure from
P to P.
143
For a real gas, the change in chemical potential with pressure is given by:
T, P ( ) T, P' ( ) = RT ln
f
f'






= V
m
real
dP
P'
P


For an ideal gas, the change in chemical potential with pressure is given by:
T, P ( ) T, P' ( ) = RT ln
P
P'






= V
m
ideal
dP
P'
P


If we subtract these two equations from each other, we get:
( )
( )

P
P
ideal
m
real
m
P
P
ideal
m
real
m
dP V V
P f
fP
RT
dP V V
P
P
RT
f
f
RT
'
'
'
'
ln
'
ln
'
ln

We then express the molar volume of the real gas in terms of the
compression factor, Z, which for a real gas is defined by Z = PV
m
real
/ RT.
We write the molar volume of the ideal gas as V
m
ideal
= RT / P. Then we take
the limit on both sides for P = 0. When the pressure P tends towards zero,
the fugacity must become close to the pressure, since the real gas at very low
pressure must behave ideally. Therefore as P approaches zero, f approaches
P and the ratio P / f approaches unity. This leads to:
( )
[ ]


= =


= =


P
P P
ideal
m
real
m
dP
P
Z
P
f
dP
P
Z
RT dP V V
P
f
RT
0
0 0
1
exp
1
ln


144
Therefore, if we know Z for a real gas, either through isothermal
measurements of the molar volume as a function of pressure or through the
knowledge of the Virial coefficients, we can calculate the fugacity
coefficient and the fugacity as a function of pressure. Once we know f(P),
we can calculate (T,P) (T,P
0
). Expressions for the chemical potential of
real gases are very important when studying chemical reactions at high
pressures, where the ideal gas law fails.

Vous aimerez peut-être aussi