Вы находитесь на странице: 1из 79

ENTROPY AND THE SECOND LAW OF THERMODYNAMICS

1. Introduction .................................................................................... 1
2. Entropy and probability ................................................................. 1
3. Relationship between entropy and properties of an ideal gas... 5
4. The Second Law of Thermodynamics ......................................... 7
5. Examples of calculations of entropy changes ............................. 7

ENTROPY AND THE SECOND LAW OF THERMODYNAMICS

1. Introduction
Our motive for studying thermodynamics is to be able to predict whether a
reaction or change in a system will occur spontaneously. Most spontaneous
reactions evolve heat (exothermic) but there are exceptions. KI dissolves
spontaneously in water but the water gets cooler and ice spontaneously melts at
25C. Both these changes are endothermic. A second factor also contributes to
the spontaneity of the system; this factor is the tendency of the system to become
more random (disordered). This factor is called entropy.

Systems tend to proceed spontaneously toward an equilibrium state. When a hot
object and a cold object are brought into contact, they proceed to a common
temperature. If a container of gas is opened to the atmosphere the gas becomes
uniformly distributed through the atmosphere. Solutes or solvents always diffuse
through a membrane in such a way that the concentrations on both sides of the
membrane become equal. Electrochemical cells always lose their ability to provide
current.

Entropy, S, is a state function which is used as a measure of capacity for change. It
originally arose as part of a discussion of the efficiency of a theoretical heat
engine. That discussion is beyond the scope of this course. In 1896, Boltzmann
explained entropy using the concept of probability and we will introduce entropy
by this means, beginning with a very simple model of two molecules being
distributed between imaginary compartments in two interconnected containers.

2. Entropy and probability
Reference: C.W. Castellan, Physical Chemistry, (Addison-Wesley, 1971) pp 195-200

Assume we have a system of two containers, interconnected so that molecules tray
move from one to the other when a tap is opened. Assume that each container has
two compartments, each of which may hold one molecule, and that initially both
of the molecules to be considered are in the one container.
i.e.



2




Now the tap is opened and the two molecules are free to move wherever they
choose and to fill the two containers in all possible ways. The resulting
distinguishable arrangements for two identical molecules are:



There are six possible arrangements. The probability of any one arrangement is
1/6.

If the number of compartments in the two containers is increased to four
1
you can
see from the diagram below that the total number of possible arrangements is 28.

The latter situation represents a state of greater disorder than the former in that
there are more possible arrangements.



The entropy, S, of a system in a specified state can be defined in terms of the
number of possible arrangements of the particles composing the system using the
equation (1), proposed by Boltzmann.

S = kln W (1)

where k = Boltzmann constant

W = probability or number of possible arrangements

1
The system consists of two containers. Each container has four compartments.

3





For example: For the first situation: For two molecules confined to the left hand
side container w = 1.

S
1
= kln 1 = 0

For two molecules to be anywhere in the two containers

W=6

S
2
= kln 6 = 1.79k

The entropy change associated with the expansion of the system from two
compartments to four compartments is

S = S
2
- S
1
= kln 6 for two molecules

=
2
k
ln 6 for one molecule

For the expansion from two compartments to four:

S = S
3
- S
1
= kln 28 for two molecules.

Notice that S is positive and increases as the number of possible arrangements
increases. A change from a less probable state for a system to a more probable
state is associated with an increase in entropy.

Now we will generalize the discussion for the case of two molecules.

If the containers have N
1
cells into which two molecules may be placed, there are
N
1
ways to place the first molecule and, for each choice of compartment for the
first molecule there are (N
1
- 1) choices for the second molecule, i.e. the total
number of arrangements is N
1
(N
1
- 1). However, since the molecules are
indistinguishable, the number of distinguishably different arrangements is:

W =
( )
2
1 N N
1 1

(2)

4




S
1
= kln [ N
1
(N
1
-1) ] (3)

If the number of compartments is increased by N
2


S
2
= kln [ N
2
(N
2
- 1)] (4) (4)

S = S
2
- S
1


= kln
( )
( )
|
|
.
|

\
|

1 N N
1 N N
1 1
2 2
(5)

If N
2
= 4, N
1
= 2 we have the situation studied above i.e. S = k ln 6.

Suppose N
1
and N
2
are very large:

then N ~ (N
1
- 1)

N
2
~ (N
2
- 1)
Then S = kln
2
1
2
N
N
|
|
.
|

\
|
(6)
= 2kln
1
2
N
N
for two molecules (7)
Now let us assume that we are dealing with molecules of an ideal gas. The position
of a molecule at any time is the result of pure chance, and since molecules are
independent, the proximity of other molecules does not affect the chance of a
molecule being where it is. We may imagine that we are dealing with imaginary
compartments of a given size, in which case the number of compartments would
be proportional to the volume available to the gas.

i.e. N
1
V
1


N
2
V
2


and
1
2
1
2
V
V
N
N
=

Thus, for two molecules we may substitute in (7)


5




S = 2kln
1
2
V
V
(8)

This treatment may be extended to the Avogadro number of molecules, provided
that the number of locations available for the molecules is very large compared to
the number of molecules. Equation (6) becomes:

S = kln
A
1
2
N
N
N
)
`

|
|
.
|

\
|


= kln
A
1
2
N
V
V
|
|
.
|

\
|


= N
A
kln
1
2
V
V
(9)
but you will remember from Section 2.4 that

k =
A
N
R


(9) becomes S = Rln
1
2
V
V
(10)

From the standpoint of a statistical definition of entropy, expansion of a gas from
a volume V
1
to a volume V
2
increases entropy because there are more ways to
arrange a given number of molecules in a larger volume than in a smaller volume.

3. Relationship between entropy and properties of an ideal gas
From the First Law of Thermodynamics:

dU = dq + dw (11)

For an ideal gas the internal energy is constant at constant temperature. Thus, for
an isothermal reversible change

dU = 0


6




and dq
rev
= -dw
rev


but we have deduced in Section 2. that

dw
rev
= - nRT ln
1
2
V
V

dq
rev
= nRT ln
1
2
V
V


For one mole,

1
2 rev
V
V
ln R
T
dq
= (12)

Substitute from (11) into (13) for an infinitesimal change

i.e. dS =
T
dq
rev
(13)

Remember that o w
rev
is the minimum amount of work done on a system.

Thus, for non-reversible process i.e. one taking place under other that equilibrium
conditions

dS >
T
dq
irrev
(14)

and in general

dS >
T
dq
(15)

Equation (15) is the defining equation for the entropy function. To assist us to
understand entropy we have chosen a very special set of conditions to derive this
function. You should understand however that equation (15) is a general equation
which may be derived in other ways which are beyond the scope of this course.


7




4. The Second Law of Thermodynamics
In this course we are particularly interested in equation (15) as a useful operational
statement of the Second Law of Thermodynamics for a system not completely
isolated from its surroundings. A more general statement is possible for a system
which is totally independent, because in such a case there will be no exchange of
heat i.e. dq
rev
= 0. Then (15) becomes

dS > 0 (16)

This equation maybe associated with the statements given on p.169 of G.M.
Barrow, Physical Chemistry, (McGraw-Hill, 1973).

The two statements that constitute the second law are:

1. When a process is carried out reversibly the entropy change in the universe of
the process is zero ..........

2. For processes that proceed irreversibly i.e. out of balance and therefore
spontaneously, the entropy of the universe of the process increases."

i.e. S
universe
> 0 (compare with the First Law U
universe
= 0).


5. Examples of calculations of entropy changes

(1) Ideal gas expansion
Calculate S for the isothermal reversible expansion of 2.00mol of an ideal gas
from 10.0 to 12.0 L at 27C.

Solution
For the isothermal reversible expansion of an ideal gas

S =
T
q
rev
= nR ln
1
2
V
V

where V
2
is the final volume


8




V
1
is the initial volume

= 2 x 8.314ln
0 . 10
0 . 12


= 3.03 JK
-1
Ans.

(2) Phase changes
Calculate the entropy change occurring when one mole of water is converted to
steam reversibly at its boiling point and 1atm pressure. (The heat of vaporization
of water = 40.67 kJ mol
-1
).

Solution
For a reversible process:

S =
T
q
rev

=
373
40670


= 109JK
-1
Ans.

(3) Chemical Reactions
Since entropy is a state function, Hess's Law also applies. Thus for a reaction such
as

2H
2
(g) + O
2
(g) 2H
2
O(l) at 25
o
C

S
reaction
= 2S
o
(H
2
O(l)) - 2S
o
(H
2
(g)) - S
o
(O
2
(g))

You will observe here that entropy content, S is given (compared to enthalpy
change, H). This is possible because we have a reference point for entropy -
absolute zero. The Third Law of Thermodynamics states a perfect crystal is
perfectly ordered at absolute zero (i.e. s = 0).

As a consequence, S for elements is not zero at 25C. Tables in Unland also list
S.

S
reaction


9





= 2(69.6) - 2(130.6) - 205

= -326J.

Since the evolution of a gas is the biggest change in entropy of the phase changes,
a useful approximation is if n
gas
> 0 then S > 0.

The emphasis in this section has been on the concept of entropy. The definition
dS= dq
rev
/T is fundamentally important but not very useful for calculation.

Entropy calculations for irreversible pathways are usually calculated by trying to
match this irreversible pathway to a series of reversible pathways. The examples in
5. are the only ones we need to consider.






























10




LIQUID VAPOUR EQUILIBRIA FOR PURE SOLVENTS AND
SOLUTIONS


1. Introduction ................................................................................. 11
2. The vapour pressure of pure liquids .......................................... 11
3. Phase diagram for a pure substance ......................................... 13
4. The vapour pressure of solutions: Raoult's Law ...................... 14
5. Worked example ........................................................................... 16
6. Henry's Law ................................................................................. 17
7. Worked example ........................................................................... 18
8. Liquid-vapour equilibria for non-ideal solutions ....................... 19
9. Pressure-composition diagrams for ideal solutions ................. 20
10. Boiling point - composition diagrams for ideal solutions ........ 21
11. Vapour-pressure composition and boiling point-composition
diagrams for non-ideal solutions ....................................................... 23
LIQUID VAPOUR EQUILIBRIA FOR PURE SOLVENTS AND
SOLUTIONS

1. Introduction
As you have seen previously, when pressure is sufficiently increased or
temperature sufficiently lowered, a real gas will condense to a liquid, still in
equilibrium with a vapour or gaseous phase. Liquids represent a condensed phase
of matter for which a given amount has a fixed volume under given temperature
and pressure conditions. Liquids, like gases and unlike solids, take the shape of
their container.

Some structure may be recognised in the liquid state, but basically the liquid is
quite a disordered arrangement.
2. The vapour pressure of pure liquids
Like molecules in the gaseous phase, molecules in a pure liquid are in constant
random motion with a distribution of velocities. Occasionally a molecule travelling
at a relatively high velocity breaks through the surface of the liquid and enters the
gaseous or vapour phase. This is the process of evaporation. Increasing the
temperature of the liquid increases the kinetic energy of the molecules and favour
evaporation. In the reverse process, when molecules of low kinetic energy in the
vapour phase collide with the liquid surface and are trapped, condensation
occurs.

If the liquid and vapour are enclosed and the space is not so large that the liquid
evaporates completely, then eventually, at a fixed temperature, an equilibrium will
be established.

liquid vapour

For a particular liquid, the pressure exerted by the vapour is a function only of the
temperature and is called the vapour pressure of the liquid at that temperature.

As temperature is increased, so is the vapour pressure. Typical data for two liquids
is shown graphically below.

(Data is from Barrow, G.M., Physical Chemistry, (McGraw-Hill, 1973, page 542.)
12


The quantity of heat absorbed in the conversion of one mole of a liquid into one
mole of a gas is the molar heat (or enthalpy) of vaporization. The more strongly
the molecules of a liquid are attracted to one another, the more heat is required to
vaporize the liquid. Thus water, in which molecules are strongly attracted by
hydrogen bonding in addition to normal van der Waal's forces, will have a much
higher enthalpy of vaporization than will, say, diethyl ether where only van der
Waal's forces are significant. Actual data are:

H
vap
(water) = 44.8kJ mol
-1
at 0C

H
vap
(diethyl ether) = 28.8kJ mol
-1
at 0C

The temperature at which the vapour pressure of the liquid equals atmospheric
pressure is the normal boiling point of the liquid. Thus the boiling point of pure
water is 100C and the boiling point of pure diethyl ether is 35C at 101,325Nm
-2

in each case.

Of course at a given temperature the vapour pressure of a liquid with high
cohesive forces between the molecules (non volatile liquid) will be lower than for
one with weak cohesive forces (volatile liquid). Again, from the diagram above it
can be seen that the vapour pressure of water at 20C is 0.023atm while the
vapour pressure of diethyl ether at 20C is 0.581atm.

You should also remember that, since the molecules moving at the highest
velocities are the ones most likely to escape from a liquid, the average kinetic
13
energy of the remainder tends to fall, so that evaporation has a cooling effect.
Evaporation of perspiration on the skin is an important mechanism for cooling
the body and the effect is even more noticeable if the evaporating liquid is more
volatile, one such as alcohol (in methylated spirits, for example).

When the escaping molecules are not trapped in contact with the liquid (as in a
vessel open to the atmosphere) equilibrium is not established, and evaporation
goes to completion.

Such processes are aided by mechanical removal of the vapour from the vicinity of
the liquid, as when a wind blows over the surface of the liquid.

Apart from increasing the temperature, the rate of evaporation may be increased
by lowering the external pressure. While the normal boiling point of a liquid is
defined as the temperature at which the liquid boils under a pressure of one
atmosphere, the liquid may be made to boil at a lower temperature simply by
lowering the external pressure e.g. by creating a vacuum above the liquid. Thus at
an external pressure of 0.100 atmospheres, water will boil at 46C.

3. Phase diagram for a pure substance
The discussion above related only to liquid vapour equilibria. As temperature and
pressure are varied for a pure substance, three phases are of significance - solid,
liquid and gaseous. A typical phase diagram relating all three is shown in the
sketch below:


TC represents the liquid-vapour equilibrium discussed in the previous sections,
while BT represents the solid-liquid equilibrium e.g. ice water and AT represents
14
the solid-vapour equilibrium e.g. ice water vapour at temperature below 0C at
one atmosphere pressure. T is known as the triple point, and represents the one set of conditions at which liquid, solid and vapour are simultaneously in equilibrium. The temperature of the triple point for water is 0.0098C.

The freezing point of the solvent is the temperature at which solid and liquid are
in equilibrium at one atmosphere pressure. For water this is 0C.

4. The vapour pressure of solutions: Raoult's Law
A solution may be defined as a homogeneous mixture of two or more
components that form a single phase. We will begin by considering so called ideal
solutions consisting of two similar liquids mixed together.

An ideal solution is one in which it is assumed that all intermolecular forces are
equal. i.e. in a solution consisting of A and B it is assumed that A-A, A-B and B-B
intermolecular forces are equal. Such a situation is closely approached by mixtures
of benzene (I) and toluene (II).



The thermodynamic consequences of ideality is that there are no enthalpy changes
associated with the mixing process. The free energy change G, is given generally
by

G = H - TS

If H = 0

Then G = -TS

Thus the free energy change associated with mixing arises only as a result of
entropy changes. It should be obvious that the mixing process will lead to an
increase in entropy for the system.

In the vapour phase above a solution there will be molecules of both species
involved i.e. each will contribute a partial pressure to the total pressure. According
to Dalton's Law, the total pressure P is given by
15

P = P
1
+ P
2


where P
1
and P
2
are the partial pressures of components 1 and 2 respectively. For
an ideal solution

P
1
= X
1
.
P
1
o


P
2
= X
2
.
P
2
o


where X
1
and X
2
are the mole fractions and P
1
o
and P
2
o
are the vapour pressures
of the pure liquids respectively. These two equations express Raoult's Law i.e. the
pressure of any component is equal to its mole fraction X multiplied by its
pressure in the pure state.

Mole fraction is defined as the number of mole of the species under
consideration divided by the total number of mole of all species in the solution.

Thus, if there are n
1
mole of component 1 and n
2
mole of component 2, then

X
1
=
2 1
1
n n
n
+


X
2
=
2 1
2
n n
n
+


and X
1
+ X
2
= 1

A plot for the benzene-toluene system which obeys Raoult's Law is shown below.
The temperature is 20C.

16


The above properties are reasonable on a qualitative basis. Since evaporation is a
surface phenomenon, and since the area of surface occupied by each type of
molecule will be proportional to its mole fraction in the solution, then the number
of molecules of each type in the vapour above the liquid should also be related to
mole fraction.
5. Worked example
The vapour pressures of pure benzene and pure toluene at 60C are 5.13 x l0
4

Nm
-2
and 1.85 x 10
4
Nm
-2
respectively. Calculate the partial pressures of benzene
and toluene, the total vapour pressure of the solution and the mole fraction of
toluene in the vapour above a solution with 0.60 mole fraction of toluene.

Solution
From Raoult's Law:

P
i
= X
i
P
i
o


where P
i
= partial pressure of component i

Xi = mole fraction of component i

P
i
o
= vapour pressure of pure i

For toluene,

17
P(toluene) = 0.60 x 1.85 x 10
4


= 1.11 x 10
4
Nm
-2
Ans.

Since X(toluene) + X(benzene) = 1.00

X(benzene) = 0.40

For benzene

P(benzene) = 0.40 x 5.13 x 10
4


= 2.05 x 10
4
Nm
-2
Ans.

From Dalton's Law of Partial Pressures

P = P(toluene) + P(benzene)

= 3.16 x 10
4
Nm
-2


Since mole fraction is proportional to partial pressure

X(toluene) =
4
4
10 x 16 . 3
10 x 11 . 1


= 0.351 Ans.

[Note that the vapour is not as rich in the less volatile component (toluene) as is
the solution.]

6. Henry's Law
When the solute in a dilute solution is volatile it is found that

P
2
= k X
2


where P
2
= vapour pressure of the solute

18
X
2
= mole fraction of the solute

k = constant dependent on the nature of the solute and
solvent and the temperature.

This relationship is known as Henry's Law.

It is also convenient to write Henry's Law, as an alternative form:

P
2
= k
m
m
2

where P
2
= vapour pressure of the solute (2)

m
2
= molality of component 2

k
m
= constant dependent on the nature of the solute and solvent
and the temperature.

Henry's Law is of importance since it is particularly applicable to the dissolution of
small concentrations of gases in liquids. The effervescence observed when a bottle
of drink is opened is an example of the decrease in the solubility of a gas as its
partial pressure is lowered. Another is the problem of dissolved nitrogen in the
blood of deep sea divers. At a point 40m below the surface of seawater the total
pressure is about 6atm. The solubility of nitrogen in plasma is then about
3.7x10
-3
mol kg
-1
, nine times that at sea level. If the diver swims upward rapidly
dissolved nitrogen gas will start boiling off and bubbles in the bloodstream will
result in dizziness and possibly death.

Henry's Law is really only applicable when there is no reaction between the gas
and the solvent. Thus it is not applicable to oxygen in blood since the oxygen
reacts with blood components such as haemoglobin.
7. Worked example
The Henry's Law constant k' for nitrogen in water at 298K is 1610atm mol
-1
kg.
What is the concentration of N
2
in water in contact with air at 298K?

Solution
From Henry's Law:


2
N
P = k'
2
N
m

19
where
2
N
P = partial pressure of nitrogen


2
N
m = molar concentration of N
2


Since air contains 78% nitrogen by volume, its partial pressure in air is 0.78atm.

k
P
m
2
2
N
N
'
=


1610
78 . 0
= 4.8 x 10
-4
mol kg
-1
of water

8. Liquid-vapour equilibria for non-ideal solutions
Raoult's Law provides a mathematical description for the behaviour of ideal
solutions i.e. those solutions in which all intermolecular forces are equal. In
general this situation does not hold.

If the force of attraction between different molecules in a solution is stronger than
between identical molecules, then the molecules of each type have more trouble in
escaping from the solution than they would from an ideal solution. Thus the
vapour pressure due to each component and the total vapour pressure will be less
than that predicted from Raoult's Law.


If the force of attraction between unlike molecules is less than between identical
molecules, then the vapour pressure due to each component and the total vapour
pressure will be higher than predicted from Raoult's Law.

Note from the diagrams that in the first case there is a minimum in the total
vapour pressure-solution composition curve (negative deviation from Raoult's
20
Law), while in the second case there is a maximum (positive deviation from
Raoult's Law).

9. Pressure-composition diagrams for ideal solutions
A vapour pressure-composition diagram is a device for showing, on one diagram,
the composition of liquid and vapour phases which are in equilibrium over the
entire range of solution composition from pure component 1 to pure component
2.

Consider an ideal solution containing components 1 and 2. From Raoult's Law:

P
1
= X
1
.
P
1
o
P
2
= X
2
.
P
2
o


where P
1
, P
2
are vapour pressures due to components 1 and 2

X
1
, X
2
are mole fractions of components 1 and 2 in the liquid phase

P
1
o
, P
2
o
are vapour pressures of pure 1 and 2.

Using Dalton's Law of Partial Pressures (pressure fraction is proportional to mole
fraction),

v
1
X =
2 1
1
P P
P
+

v
2
X =
2 1
2
P P
P
+


where
v
1
X and
v
2
X are the mole fractions of components 1 and 2 in the vapour
phase. These equations may be rewritten:

v
1
X =
2 1
o
1 1
P P
P X
+

v
2
X =
2 1
o
2 2
P P
P X
+



The ratio of
v
1
X to
v
2
X is

21
V
2
V
1
X
X
=
O
2
O
1
2
1
P
P
X
X


This equation may be used qualitatively to show that the vapour above a solution
of composition (X
1
, X
2
) will be relatively richer in component 1

(
V
1
X > X
1
) and poorer in component 2 (
V
2
X < X
2
) if
O
1
P is greater than
O
2
P .

Study Example 9.6 again and notice that you have actually followed the above
derivation there.

The equation may be used to calculate points for a vapour pressure-composition
diagram such as that shown below, once
O
1
P and
O
2
P are known.



In such a diagram, the X axis is used to represent both liquid and vapour
composition. Thus the solution corresponding to point A on the liquid line has a
vapour pressure P and the composition of the vapour above the liquid may be
found by extrapolating horizontally to the vapour line at point B and down to the
X axis. Note again that
V
2
X is greater than X
2
since component 2 is the most
volatile of the pair. Remember that the vapour line is always under the liquid line.

10. Boiling point - composition diagrams for ideal solutions
Since laboratory processes are more often performed at constant pressure (one
atmosphere) than at constant temperature, diagrams which show the relationship
between liquid composition, vapour composition, and temperature are of
22
considerable practical significance. Such diagrams are difficult to deduce
mathematically, and are normally based on experimental data.

Again, using qualitative reasoning, we may see that the pure liquid with
the highest pressure will have the lowest boiling point, and vice versa. At constant
pressure, the liquid is the stable phase at low temperature, so the liquid line is now
below the vapour line.


Allow a liquid of composition X
2
to evaporate at constant pressure. When the
vapour above the liquid is condensed it has a composition X, which is richer in the
more volatile component 2 than is the original liquid. If the liquid X were
collected and re-evaporated, it would be still richer in the more volatile
component (composition X
1
) .

This discussion introduces the subject of distillation - the separation of liquids of
differing volatility by heating the solution and collecting and condensing the
vapour which is richer in the more volatile component. By repeated application of
the technique it is possible to separate completely mixtures whose behaviour is
described by the graph above.

In practice, the process of fractional distillation combines many hypothetical
distillations into a single process. The type of apparatus used is shown in the
diagram below.

23

The high boiling fraction, rich in the least volatile component, is concentrated in
the pot, while the most volatile components concentrate at the collection point at
the top of the still. A gradient is established in between.

Fractional distillation is widely employed in industry and fractionating columns are
a particularly important part of any plant such as an oil refinery. Here the
petroleum is separated into fractions such as gasoline (boiling range 40 - 200C),
kerosene (boiling range 175 - 325C), etc.

11. Vapour-pressure composition and boiling point-composition
diagrams for non-ideal solutions
In general solutions are not ideal. The most commonly encountered situation has
a maximum in the vapour pressure-composition diagram and consequently a
minimum in the boiling point-composition diagram. The diagrams are complex
and are normally determined experimentally. A consequence of the existence of
minima or maxima in the curves is that at some point vapour and liquid in
equilibrium have the same composition, and it is not possible to separate
completely the two components of the mixture. Study the diagram below which is
typical of those for mixtures with a minimum boiling point.
24

Begin with a solution containing X
1
mole fraction of component 1. If this liquid is
heated it will begin to boil at t
1
and if a sample of the vapour is collected and
condensed, the condensate will have composition X
2
. If a solution containing X
3

mole fraction of component 1 is heated it will begin to boil at t
2
and the collected
distillate will have composition X
4
. Note that in each case, repeated distillation
processes will lead to a distillate which has the composition X corresponding to
the minimum point in the curve, while the solution in the pot becomes richer in
the pure component 2 in the first case (X
l
to X
2
) and pure component 1 in the
second case (X
3
to X
4
).

If one started with a solution of composition X, the vapour in equilibrium will
have identical composition, so that, at a given pressure, separation of such a
solution into its components is not possible. Such a solution is called an
azeotrope. An azeotrope mixture distills unchanged at a given pressure (the
vapour has the same composition as the liquid).

If a solution with a composition lying to the left of the minimum in the above
diagram is distilled so that the distillate having the azeotropic composition is
collected, the mixture remaining in the pot will become richer and richer in
component 2. Ultimately, the point may be reached where only component 2
remains in the pot. This component may then be recovered in a pure state. From a
solution with a composition lying to the right of the minimum in the diagram,
pure component 1 may ultimately be isolated from the pot.

One of the best known examples of the above behaviour is the ethanol-water
system. At normal atmospheric pressure, water boils at 100.0C and ethanol boils
at 78.5C. Ethanol and water form an azeotrope of boiling point 78.2C with a
composition of 95.6% ethanol and 4.4% water by weight. Fermentation results in
ethanol/water solutions containing, say 10% ethanol. Distillation is used to
increase the ethanol concentration, but instead of yielding 100% ethanol, 95%
ethanol is obtained. For many purposes this mixture is used directly.
25


26
NON-IDEAL BEHAVIOUR OF GASES

1. Introduction .................................................................................. 27
2. Condensation of gases and the critical point ........................... 27
3.1 Introduction 29
3.2 Correction for excluded volume 29
3.3. Correction for attractive forces between molecules 29
NON-IDEAL BEHAVIOUR OF GASES
1. Introduction
Actual gases exhibit, to some extent, deviations from the behaviour predicted by
the ideal gas equation of state. These deviations are particularly important at

(i) high pressures
(ii) low temperatures

To show deviations from ideality it is convenient to use a factor z called the
compressibility factor. With this factor

PV = znRT (1)

and z =
nRT
PV
(2)

For one mole of an ideal gas, z must have the value one. Examine Figure 2 which
shows plots of z versus pressure for hydrogen and for methane at two different
temperatures. Notice that the gases show behaviour close to ideal at pressures
close to 1atm and even at 100atm, the deviations for hydrogen at 0C and
methane at 100C are about 7% and 3% respectively. Notice also that the
deviations increase as pressure increases and as temperature decreases.

2. Condensation of gases and the critical point
If there were really no attractive forces between gas molecules, then gases would
never condense to liquids. They would always completely fill the containing vessel.

We are able to observe the onset of non-ideal behaviour by studying
pressure-volume isotherms of the type introduced in the discussion on Boyle's
Law. Such isotherms are shown in Figure 3.

At high temperatures the isotherms show only slight deviations from ideal gas
behaviour. They also conform to ideal behaviour at low pressures and large
volumes. However, at low temperatures very marked deviations occur. Follow one
such isotherm. From A to B, as pressure is increased volume decreases
approximately in accordance with Boyle's Law. At point B the gas begins to
condense to a liquid and the pressure remains constant at the equilibrium vapour
pressure for the liquid at that temperature i.e. any attempt to increase the external
pressure simply results in the condensation of more liquid without change in
28

pressure. By point C all gas has condensed to liquid and the pressure may again
increase. However, because it is very difficult to compress a liquid, very large
increases in pressure result in only minute volume changes i.e. the isotherm is
almost vertical.

Figure 2: Plot of compressibility factor z for one mole of gas, where z =
nRT
Pv


Note that the diagram shows the pressure-volume conditions over which the
various phases are present.

Figure 3: Typical PV isotherms for a real gas in the region of the critical point (T l > T2 > T3)

29

One isotherm is of particular significance. This is the so-called CRITICAL
ISOTHERM which is the one which just shows a point of horizontal inflection
at the top of the zone which indicates equilibrium between the gas and liquid
phases. The point of horizontal inflection is called the CRITICAL POINT.

The temperature corresponding to the critical isotherm is known as the
CRITICAL TEMPERATURE. The particular significance of this temperature
is that it represents the highest temperature at which the gas may be condensed by
the application of pressure alone.

CRITICAL PRESSURE and CRITICAL VOLUME are the pressure and
volume per mole respectively at the critical point.

3.1 Introduction
In 1873, Van der Waals attributed the failure of the ideal gas equation to duplicate
the behaviour of real gases to the neglect of

(i) Volume occupied by gas molecules
(ii) Attractive forces between the molecules

3.2 Correction for excluded volume
When n mole of gas is placed in a container of volume V, the volume in which the
molecules are free to move is equal to V only if the volume of the gas molecules
themselves is negligible. The presence of molecules of finite size means that a
certain volume - called the excluded volume is not available for movement by the
molecules.

If the volume excluded by one mole of molecules is represented by b, then
P(V-nb)=nRT is a better equation than PV = nRT for the description of the
behaviour of real gases, since it gives the real free volume for movement by
molecules. Values of b must be measured empirically for each gas.

Note that nb becomes more significant for a given amount of gas as V decreases
i.e. as P increases. This is in accord with the experimental fact that deviations from
ideality are most serious at high pressure.

3.3. Correction for attractive forces between molecules
The presence of attractive forces between molecules is indicated by the fact that
all gases condense to liquids at low enough temperatures i.e. attractive forces
30

between molecules are sufficiently large to overcome thermal motion due to
kinetic energy. This motion normally keeps a gas filling its container rather than
settling in bulk under the influence of gravity. The nature of Van der Waals forces
has been studied in Inorganic Chemistry I.

The Van der Waals attractive forces act with the pressure to hold molecules
together. Van der Waals saw these attractive forces as being particularly significant
as a molecule was about to strike the walls of the container and contribute to the
pressure. Attractive forces from within the body of the gas pull the molecule back
from the wall, reducing pressure. These forces act throughout the body of the gas
and their effect is related to gas density expressed as
V
n
. Since there is a two way
interaction between gas molecules, Van der Waals made the correction term
proportional to
2
V
n
|
.
|

\
|
with a constant of proportionality i.e. gas pressure, or the
total pulling together of molecules is increased by
2
2
V
n a
by Van der Waal's forces.
The Van der Waal's equation of state for a real gas then becomes:

|
|
.
|

\
|
+
2
2
V
an
P (v - nb) = nRT

These attractive forces will be most important when

(i) molecules are forced into close proximity i.e. at high pressure
(ii) when molecules are moving along slowly and spend long periods adjacent to
one another i.e. at low temperature.

Example:
Heald, C., Smith, A.C.K., Applied Physical Chemistry, (MacMillan, 1974), p.13.

An autoclave of 1.00L capacity is filled with 644g of carbon tetrachloride and then
heated to 650K. It is desired to calculate the pressure developed. The Van der
Waals coefficients for carbon tetrachloride are

a = 2.04Nm
4
mol
-2
and b = 1.38 x 10
-4
m
3
mol
-1
; the molar mass of this substance
is 154g.

31

Solution
number of mole of CCl
4
=
154
654
= 4.18

Van der Waals equation is:

|
|
.
|

\
|
+
2
2
V
an
P (V - nb) = nRT
Substitute the values given.
( )
|
|
.
|

\
|
+

2
3
2
10 x 00 . 1
18 . 4 x 04 . 2
P x (1.00x10
-3
- 4.18x1.38x10
-4
)

= 8.314 x 4.18 x 650

i.e. (P + 3.56 x 10
7
) 4.23 x 10
-4
= 2.26 x 10
4


P = 17.9MNm
-2
(177atm).

Calculate the pressure using the ideal gas equation and compare the two.























32


THE GIBBS FREE ENERGY FUNCTION

1. Introduction .................................................................................. 33
2. The Gibbs free energy function ................................................... 33
3. Worked example ........................................................................... 35
4. Standard free energies ................................................................ 35
5. Relationship between free energy and pressure ....................... 37
6. Relationship between standard free energy change and the
equilibrium constant ........................................................................... 39
7. Worked example .......................................................................... 40
8. Temperature dependence of free energy and equilibrium
constant ............................................................................................... 44
9. Le Chatelier's principle ............................................................... 45
10. Worked example .......................................................................... 46
THE GIBBS FREE ENERGY FUNCTION

1. Introduction
We now have a criterion for spontaneous reactions viz S
universe
> 0 (S
universe
= 0
for reversible processes). However, entropy changes for the whole universe are
difficult to measure. Note that for reactions which proceed spontaneously but
S
system
< 0 then there must be an entropy increase of equal or greater magnitude
in the surroundings. When water is frozen in a fridge S < 0 but the motor gives
off heat, causing an increase in the entropy of the air molecules.

The tendency for chemical reactions to move towards equilibrium is influenced by
two factors

(i) the tendency towards minimum energy
(ii) the tendency towards maximum entropy.

A new state function, called Gibbs Free Energy (G) considers both the energy
(enthalpy) and entropy components of a change (at constant pressure and
temperature.)

2. The Gibbs free energy function
From the Second Law, for any reaction at constant T

S
2
- S
1
> q/T (1)

From the First Law

U
2
- U
1
= q + w (2)

S
2
- S
1
>
T
w U U
1 2

(3)

if only pressure-volume work is considered

S
2
- S
1
>
( )
T
V V P U U
1 2 1 2
+
(constant T, P) (4)

(Since w = -PV at constant P)
34

S
2
- S
1
>
( ) ( )
T
PV U PV U
1 1 2 2
+ +
(5) (5)

but H = U + PV

S
2
- S
1
>
T
H H
1 2

(6)
for a spontaneous charge

T(S
2
- S
1
) - (H
2
- H
1
) > 0 (7)

or (H
2
- H
1
) - T(S
2
- S
1
) s 0 (8)

Define G by G = H - TS

at constant T, G = H - TS

(8) becomes

G < 0 spontaneous
G = 0 reversible
G > 0 no spontaneous reaction

Thus for a reaction to proceed spontaneously towards equilibrium in the direction
written G must be negative. A reaction for which G is positive will not proceed
spontaneously as written. In fact it should proceed spontaneously in the reverse of
the direction being considered.

It must be noted here that thermodynamics says nothing about the speed with
which a reaction will occur. Thus, while a spontaneous reaction may be predicted
by the G value, the reaction may be so slow that there is no evidence of its
occurrence.

Factors influencing rates of reactions will be studied in the Chemical Kinetics
course.

You can see that G is a state function because U, P, V, T and S are all state
functions.
35

3. Worked example
Barrow, G.M., Physical Chemistry, (McGraw-Hill, 1973) page 193. Calculate the free
energy change for the process of converting 1mol of water at 100C and 1atm to
steam at the same temperature and pressure.

Solution
Gibbs free energy G, is defined as

G = H-TS

For a change at constant pressure and temperature

G = H - TS

Since the process is carried out at constant pressure, the enthalpy change is equal
to the heat absorbed

i.e. H = q
p


If the change is carried out reversibly

S =
T
q
T
q
p
rev
=

G =H-TS

= q
p
- T
T
q
p


= 0

This means that since water and steam are in equilibrium at 100C, the free energy
change for the conversion of one into the other must be zero (i.e. the process is
reversible).

4. Standard free energies
* Tables of Standard Free Energies are to be found in Unland.
36

As a reference point, zero values of free energy are assigned to the stable form of
the elements at one atmosphere pressure. These, and the free energies of
compounds based on these references, are called standard free energies of
formation. These free energies may be manipulated in the same way as standard
enthalpies of formation (usually quoted at 25C).

Example
Barrow, G.M., Physical Chemistry, (McGraw-Hill, 1973), page 196. Calculate the free
energy change for the reaction

H
2
C= CH
2
(g) + H
2
(g) CH
3
CH
3
(g)

Does the calculated free energy change indicate that products are favoured over
reactants at standard condition?

Solution
From the tables we find that

G
f
o
(H
2
C = CH
2
(g)] = + 68.12kJ mol
-1

G
f
o
[H
2
(g)] = 0
G
f
o
[CH
3
CH
3
(g)] = -32.89kJ mol
-1


For the reaction as written

G
o
298
= G
f
o
[Products] - G
f
o
[Reactants]

= -32.89 - 68.12

= -101.01kJ mol Ans.

Since G
o
298
is negative, products are favoured over reactants at standard
conditions.

37
5. Relationship between free energy and pressure
Gibbs standard free energy data has provided us with criteria for assessing the
possibility of a reaction occurring at 25C when all reagents are in their standard
states. For the data to be of real use it is necessary to be able to make calculation
at other temperatures and pressures etc. We shall now consider the relationship
between free energy and pressure.

From the definition of free energy:

G = U+PV-TS (13)

Differentiate (1)

dG = dU + PdV + VdP - TdS - SdT (14)

From the First Law of Thermodynamics

dU = dq + dw (15)

From the Second Law of Thermodynamics

dS =
T
dq
rev
(reversible process) (16)

From the definition of reversible PV work

dw = -PdV (17)

Combining (15), (16) and (17) gives

dU = TdS - PdV (18)

dU - TdS + PdV = 0 (19)

Substitute (19) into (14)

dG = VdP - SdT (20)

38
at constant temperature

T
dP
dG
(

= V (21)

To study the change in free energy for an ideal gas we may write:

}
2
1
G
G
dG =
}
2
1
P
P
V dP at constant T (22)

For an ideal gas:

V =
P
nRT
(23)

Substitute (23) into (22)


}
2
1
G
G
dG = nRT
}
2
1
P
P
P
dP
(24)

[g]
1
2
G
G
= nRT [nRT]
1
2
P
P
(25)


G
2
- G
1
= nRT l n
1
2
P
P
(26)

If we let state 1 be the standard state and state 2 be the state under consideration
then

G
2
= G

G
1
= G

P
2
= P

P
1
= 1 atmosphere

Then G = G + RT ln P (27)
39

Notice that, contrary to our usual practice, pressure is specified in atmospheres in
equation (27) rather than in Nm
-2
.

The full derivation of (26) and (27) has been included here for reference only. The
result, however, is needed to derive the very important relationship in 6.

6. Relationship between standard free energy change and the
equilibrium constant
Consider a reaction involving four ideal gases, A, B, C and D all at temperature T:

a A + b B c C + d D

where a, b, c and d are the numbers of moles of each reagent involved.

Free energy of a mol of A = aG
A
= aG
A
o
+ aRT ln P
A


Free energy of b mol of B = bG
B
= bG
B
o
+ bRT 1n P
B

Free energy of c mol of C = cG
C
= cG
C
o
+ CRT ln P
C

Free energy of d mol of D = dG
D
= dG
D
o
+ dRT ln P
D

where G
A
, G
B
, ..., G
A
o
, G
B
o
, ... are free energies of one mol of reagent A, B, etc.

The free energy change for the reaction is given by

G = G
products
G
reactants


= cG
C
+ dG
D
- aG
A
- bG
B


= cG
C
o
+ dG
D
o
- aG
A
o
- bG
B
o
+ RT ln
( ) ( )
( ) ( )
b
B
a
A
d
D
c
C
P P
P P

or G = G
o
+ RT ln
( ) ( )
( ) ( )
(

b
B
a
A
d
D
c
C
P P
P P
at constant T

40
=G
o
+ RT ln Q (where Q is the ratio of products to reactants)

In the case where the PA , PB , PC and PD values are partial pressures of the
reagents at equilibrium then G = 0 so that

G
o
= -RT ln
( ) ( )
( ) ( )
(

b
B
a
A
d
D
c
C
P P
P P
= -RT ln Q (28)

Since G
o
for a particular reaction at a fixed temperature is constant, then the
term on the right hand side must have a constant value that is independent of the
individual pressures. We write:

K
p
=
( ) ( )
( ) ( )
(

b
B
a
A
d
D
c
C
P P
P P
(29)


and G
o
= -RT ln K
p
(30)


(Q = K
p
only at equilibrium)

where K
p
is the equilibrium constant expressed in terms of partial pressures of the
reagents involved in the reaction.

A similar result applies for reactions in solution i.e. G
o
= -RT ln K in general.

Remember that to use equation (30) pressures must be in atmospheres.

7. Worked example

1. Use tabulated standard free energies to decide whether or not it is worth
seeking a catalyst to increase the rate of production of NO(g) from a mixture
of N
2
(g) and O
2
(g) at 298K and 1atm pressure.

The reaction is:

N
2
(g) + O
2
(g) 2 NO (g)

41
From tables

G
f
o
[N
2
(g) ] = 0.0kJ mol
-1


G
f
o
[O
2
(g)] = 0.0kJ mol
-1


G
f
o
[NO (g)] = 86.69kJ mol
-1


For reaction as written

G
f
o
= 2G
f
o
[NO(g)] - G
f
o
[N
2
(g)] - G
f
o
[O
2
(g)]

= 2 x 86.69 - 0 - 0

= 173.38kJ

Since G
o
is +173.4kJ the reaction as written will not proceed to a significant
extent under standard conditions. Since a catalyst changes only the rate of
reaction and not the position of equilibrium, there is no point in seeking a
catalyst.

2. Goates, J.R., Ott, J.B., Chemical Thermodynamics, (Harcourt Brace Jovanovich,
1971), page 75.

For the reaction

2 CO (g) + O
2
(g) = 2 CO
2
(g)

write the equilibrium constant expression and calculate the value for K
p
at
25C.

Solution
At equilibrium,

G
o
= -RT ln K
p
(1)

Where

K
p
=
2 2
2
O
2
CO
2
CO
P P
P
(2)

42
From (1)

ln K
p
= -
RT
G
o
A
(3)

From tables

G
f
o
[CO (g)] = -137kJ mol
-1


G
f
o
[O
2
(g)] = 0

G
f
o
[CO
2
(g)] = -394kJ mol
-1


For the reaction

G
f
o
= 2(-394) - 2(-137) - 0

= -514kJ

Substitute in (3)

ln K
p
= -
298 x 314 . 8
1000 x 514


= 207

K
p
= 7.92 x 10
89
atm
-1 -


i.e. at equilibrium the reaction is heavily in favour of the products i.e. CO
2
.

3. Calculate G
o
and K
p
for the following reaction at 298K.

N
2
(g) + 3H
2
(g) 2 NH
3
(g).

Solution

From tables
G
f
o
[N
2
(g) ] = 0

-
This is a very large number. Some calculators may give an overflow or error message if they are asked to evaluate such large numbers.
43
G
f
o
[H
2
(g)] = 0
O
f
o
[NH
3
(g)] = -16630J mol
-1

For the reaction as written

o
298
G A = 2 (-16630) - 0 - 3 x 0

= -33260J Ans.

At equilibrium,

o
298
G A = -RT ln K
p


ln K
p
= -
298 x 314 . 8
33260


= 13.42

K
p
= 6.7 x 10
5
at 298K Ans.

(This is a very important reaction since it is the basis for the commercial
production of ammonia.)

4. A mixture of N
2
(g), H
2
(g) and NH
3
(g) is introduced into a container in
amounts which gave the following partial pressures at 298K.

P(N
2
) = 0.10atm

P(H
2
) = 0.10atm

P(NH
3
) = 10atm

Comment on what is likely to happen in such a mixture, using the data in the
example above. Calculate K
p
for the reaction.

Solution
In general G = G
o
+ RT ln Q

G = G
o
+RT ln
3
H N
2
NH
2 2
3
P x P
P


44
= -33260 + 8.314 x 298 x l n
3
2
10 . 0 x 10 . 0
10


= -33260 + 8.314 x 298 x l n(1.00 x 10
6
)

= 969J.

ln K
p
= - G/RT = 33260/8.314 x 298 =13.4

K
p
= e
13.4
= 6.6 x 10
5


Note that K is not equal to Q.

From this result we see that such a mixture is not at equilibrium, and would
proceed towards equilibrium in a direction which is the reverse of that written
above i.e. ammonia would decompose to give nitrogen and hydrogen, despite
G
o
being less than zero. It is G, not G
o
, which predicts the course of a
reaction. In this reaction G is positive and will decrease as ammonia
decomposes, until G=0 (at equilibrium) and Q = K.

8. Temperature dependence of free energy and equilibrium
constant
The quantitative relationship between K and T is easy to derive if we assume H
is independent of T (a reasonable assumption provided the temperature range is
not too great).

G
o
= H
o
- TS
o
- RT lnK (1)

at two temperatures T
1
and T
2


ln K
2
=
R
S
RT
H
o
2
o
A
+
A
(2)

ln K
I
=
R
S
RT
H
o
1
o
A
+
A
(3)

Subtracting (3) from (2)

ln
1
2
K
K
=
|
|
.
|

\
|

A
1
2
o
T
1
T
1
R
H
(4)
45

(assuming S
o
is also independent of T)

or ln K =
RT
H A
+ constant (5)

It is clear from (4) that the rate of change of equilibrium constant with
temperature depends on the standard enthalpy of the reaction. From (4) it may be
shown that when a reaction is exothermic (H < 0), the equilibrium constant
decreases as temperature increases.

9. Le Chatelier's principle
A very valuable qualitative principle for discussing the effect of pressure and
temperature changes on the position of chemical equilibrium is Le Chatelier's
Principle which may be stated:

"If a change occurs in one of the factors, such as temperature or pressure, under
which a system is in equilibrium, the system will tend to adjust itself so as to annul,
as far as, possible, the effect of that change."

(Glasstone, S., "Textbook of Physical Chemistry", (Macmillan, 1962), page 831).

Changing the pressure is of great significance only in reactions involving one or
more gaseous species, and there only when the number of mole of gaseous
reactants is different from the number of mole of gaseous products.

Consider the reaction:

N
2
(g) + 3 H
2
(g) 2 NH
3
(g); H
298
= -92.4kJ

There are four mole of gaseous reactants and two mole of gaseous products. If the
reaction is at equilibrium and the pressure is increased, the system will move to
counteract the stress by reducing the number of molecules, i.e. N
2
will react with 3
H
2
to form 2 NH
3
. Thus the forward reaction is favoured by increasing pressure
i.e. K
p
will increase.

We have seen in the previous section that an increase in temperature for an
exothermic reaction results in a reduction of the equilibrium constant. Thus
46
ammonia would decompose to nitrogen and hydrogen if temperature were
increased in the above case.

The third type of stress which is of interest is a change in the concentration of one
species. Thus if more N
2
(g) is added to the above system at equilibrium, the
system will respond by moving in the forward direction i.e. N
2
(g) and H
2
(g) will
combine to give NH
3
(g). We have met this effect in Example 3 in Section 8.

10. Worked example
In a study of the reaction

H
2
(g) + I
2
(g) 2 HI(g)

it was found that the average value of H over the temperature range 430 - 470C
was - 12.5kJ.

(i) Predict the effect on the equilibrium position of an increase in temperature.
(ii) Predict the effect on the equilibrium position of an increase in pressure.
(iii) Predict the effect of the addition of HI to the mixture.

Solution
(i) Since this is an exothermic reaction (i.e. heat is given out) and increase in
temperature will decrease the equilibrium constant i.e. HI(g) will decompose.
(ii) Since there are equal number of moles of gaseous reactants and products,
pressure changes will not effect the position of equilibrium.
(iii) Addition of HI imposes a stress on the reaction. This stress may be relieved if
HI decomposes to give H
2
(g) and I
2
(g).














47













































48
THERMODYNAMICS - THE FIRST LAW

1. Introduction .................................................................................. 49
2. Thermodynamic terms ................................................................. 49
3. State functions ............................................................................. 50
4. Pressure-volume work ................................................................. 50
5. Heat ............................................................................................... 61
5.1 Heat capacity 62
6. Summary ....................................................................................... 63
7. The First Law of Thermodynamics.............................................. 63
8. Application of the First Law under special conditions .............. 65
8.1 Processes occurring at constant volume 65
8.2 Processes occurring at constant pressure 66
8.3 Relationship between q
P
and q
V
69
9. Heat capacities at constant pressure and constant volume ..... 71
10. Relationship between C
p
and C
v
for an ideal gas ....................... 72
11. Internal energy for an ideal gas ................................................... 74
49
THERMODYNAMICS - THE FIRST LAW

1. Introduction
Thermodynamics is concerned with studying the energy changes associated with
chemical and physical systems. The energy changes associated with chemical
reactions can be used to predict the equilibrium position of a chemical reaction. Thus
the principles of thermodynamics can be used to associate the thermal properties of
a system with the equilibrium state of a chemical system. Two points need to be
made initially about thermodynamics.

(i) it is concerned only with the macroscopic properties of a system - not with the
behaviour of individual atoms or molecules.

(ii) thermodynamics tells nothing about the rate of attaining equilibrium (this will be
dealt with in the kinetics section).

2. Thermodynamic terms
Thermodynamics is concerned with systems, separated from the surroundings by a
boundary. The system together with the surroundings constitute the universe. The
positioning of the boundary is quite arbitrary (it need not be a physical boundary).



The system is open when mass passes across the boundary; closed when no mass
passes across the boundary. The system is isolated when the boundary prevents any
interaction with the surroundings.

Energy may be transferred between a system and its surroundings by:
50

(i) performance of work
(ii) exchange of heat

3. State functions
A state function is one which is used to describe the state of a system. It has the
mathematical property that the change in the function depends only on the initial
and final state of the system and not on how the change was accomplished i.e. it is
path independent.

Example
The simplest chemical system to study is the one dealt with in sections 1 - 3 i.e. an
ideal gas. The state of the gas is specified by the state functions P, V, n, T.
However, since they are related; by the Equation of State (PV = nRT), only three of
these functions need to be specified to completely define the state of the gas.

Pressure, for example, is a state function since the change in pressure, P, going
from pressure P
1
to P
2
is simply P
2
- P
1
, regardless of how the pressure was changed.

The concept of state functions should become clearer when we consider two
non-state functions (or path dependent functions), work and heat.

4. Pressure-volume work
Work is defined as the product of a force and the distance through which the force
operates i.e. w = fdx

where w = work
f = force
dx = displacement

In the SI system of units, work, like energy, is measured in joules (J)

i.e. w = f x = N x m = J

51
From our point of view we are particularly interested in pressure-volume work as is
achieved when, say, a piston is displaced by an expanding gas. Note that the product
of pressure and volume has the units of J.

P x V = Nm
-2
x m
3

= Nm
= J

Consider the work associated with the expansion or contraction of fluids. In the
following diagrams a gas is enclosed in a chamber by a frictionless, weightless piston.
The gas exerts a pressure P on the face of the piston and to keep the piston in place
an external pressure P
e
equal to P must be applied.




Note that the system is the gas in the piston and the boundary is the walls of the
piston.

If P
e
is reduced by P, the gas inside the chamber will expand and the piston will
move until the pressure of the gas (P-P) again equals the external pressure (PC-AP).
During this process a displacement x occurs in the position of the piston.

Consider a minute (infinitesimal) change: Since, by definition,

Pressure = force/unit area
52

i.e. P
e
= f / A (1)

.
.
. f = P
e
A (2)

where f is the force exerted on the piston by P
e
and A is the area of the face of the
piston.

Since w = f dx (3)

Substitute in (2)

w = P
e
Adx (4)

At this stage we introduce a sign convention which states that work is positive
when done on a system and negative when done by a system. An expanding gas is
capable of doing work, so that w in this case should be negative

w = - P
e
dV (5)

where Adx = dV, the change in volume.

Now we are concerned with the calculation of work during an entire change and not
just during an infinitesimal change, so we shall spend a little time on mathematics.

If the pressure remains constant during an expansion from V
1
to V
2
, this
may be shown graphically as:


53



The magnitude of the work is given by the product of pressure and change in
volume. This is simply given by the shaded area of the graph under the line
corresponding to P
e
and between the limits of V
1
and V
2
. In this case

w = - P(V
2
- V
1
) (6)

Mathematically, the area under a curve is obtained by integration between the limits
of V
1
and V
2
. In this case

dV P w
e
V
V
2
1
}
= (7)

Since. P
e
is constant, it may be taken outside the integral. Then (7) becomes

dV P w
2
1
V
V
e
}
=
| |
1
2
e
V
V
V P =
( )
1 2 e
V V P = (8) (8)

The above example could be expressed in an alternative way. Suppose the piston had
a block, x, resting on it. When the block is removed the gas expands until gas - P
e

(the external pressure) = P
2



54


w = - P
e
(V
2
- V
1
)

= - P
2
(V
2
- V
1
)




Suppose the change from state 1 to state 2 is now carried out a different way. Instead
of one block x, we have two blocks, each half the weight of x anti these are removed
one at a time.

55





Total work = work (1 1') + work (1 2) since both steps are constant pressure
steps

w = - P
1
' (V
1
' - V
1
) - P
2
(V
2
- V
1
' )



Note that the shaded area is less than the first case, although the initial and final
states are the same.

We could continue to divide the blocks and remove them one at a time (increasing
the work done) until we reach the limiting case - approximated below.


56

Here x is a pile of sand and we remove the grains one at a time.

In this case, work is still given directly by the area under the P-V curve. We may see
that this is so by dividing the area into a number of rectangles, each of which is
similar to the rectangle in the previous diagram. It is obviously possible to assume
that the pressure changes by small jumps from P
1
to P
2
to P
3
, etc., and if all volume
changes are made equal at V then

w = -(P
1
V+P
2
V+P
3
V+..........)

V
P
i
i
A
|
.
|

\
|E
= (9)

The estimate of the area under the curve will improve as V is made smaller. Again,
the true area is found by integration between the limits of V
1
and V
2
. To make this
possible it is necessary to know the relationship between P and V. For an ideal gas
this is obtained from the equation of state.

57

i.e. PV = nRT


V
nRT
P =

Since PdV
}
=
2
1
V
V
w

substitute for P from the equation of state for an ideal gas:

dV
V
nRT
}
=
2
1
V
V


For a given amount of gas at constant temperature

V
dV
nRT
}
=
2
1
V
V
w

| |
1
2
V
V
V ln nRT =
1
2
V
V
nRTln = (10)


(The process of integration will be quite familiar to many of you and something of a
mystery to others. For the latter group, just think of integration as a mathematical
device to find the area under a curve between limits, as shown here. The integrals we
have used here viz.

| | ( )
1 2
1
2
a
a
a a
a
a
x dx
2
1
= =
}


and | |
1
2
1
2
a
a
a
a
ln
a
a
x ln
x
dx
2
1
= =
}


58
where ln refers to natural logarithm or log to the base e, will keep on recurring,
but will be the only ones necessary for this course.

The limiting case described above is the constant temperature, or isothermal, case
and represents the maximum work done by the system. This case is a reversible
one, as the system is never more than an infinitesimal distance from equilibrium. The
system can be returned to equilibrium in a ideal reversible case with no observable
change in the surroundings. Obviously reversibility is an ideal case which can at best
only be approximated. However, for an engine, say, it does represent the best
operating conditions thermodynamically (in terms of energy flow). There is the
disadvantage that an ideal reversible system changes infinitely slowly (quite a
drawback for an engine!) but remember thermodynamics says nothing about rates.

The above case demonstrates work is not a state function. The value varies
depending on the path, and is different in the three cases (although the initial and
final states are the same). This can be illustrated in the following case.

Example
0.4 moles of an ideal gas initially at 10 atmospheres and one litre expands to a final
volume of 10L and one atmosphere;

(i) at constant pressure
(ii) in two steps; initially to a volume of 5L and 2atm
(iii) isothermally and resersibly (at 25C).

Calculate the work done in each case.

(i) At constant pressure:

P
1
= 10atm = 1,013,250 Pa

V
1
= 1L = 10
-3
m
3
(remember SI units!)

P
2
= 1atm = 101,325 Pa

V
2
= 10L = 10
-2
m
3

59

w = - P
2
(V
2
- V
1
) = -101,325 (0.01 - .001)

= - 912J (negative, since the system is doing work)


(ii) P
1
= 1,013,250
1
P = 202,650

V
1
= 10
-3
m
3
V
1
= 5 x 10
-3
m
3


P
2
= 101,325 Pa

V
2
= 10
-2
m
3


w = - 202,640(.005 - .001) - 101,325(.01 - .005)

= - 811 - 507 = -1318J

(iii) in the isothermal case equation 10 is used

w = - nRT ln V
2
/V
1

= - 0.4 x 8.314 x 298 l n
3
3
10 x 1
10 x 10



= - 2282J

A third type of process which is of interest is that which occurs at constant volume.
In this case there is obviously no change in volume, so there is no work being done
on or by the system.

i.e. w = PdV
}

2
1
V
V

= 0 since dV = 0

In summary, we are concerned with processes occurring under three different
conditions:

At constant pressure

60
w = PdV
2
1
V
V
}

= -P(V
2
- V
1
) (8) (8)

At constant temperature

w = PdV
2
1
V
V
}


= -nRTln
2
1
V
V
(ideal gas) (10)

At constant volume

( ) 0 dV since 0
PdV w
2
1
V
V
= =
=
}


Example 1
Goatcs, J.R., Ott, J.B., Chemical Thermodynamics, (Harcourt, Brace, Jovanovich, 1971),
p.13.

Calculate the work in joules involved in the constant pressure expansion of 3 moles
of gas from an initial volume of 2.00dm
3
to a final volume of 10.00dm
3
. The pressure
remains constant at 7.00atm.

Solution

( )
( )
. Ans J 5674
10 x 00 . 1 x 00 . 2 00 . 10 101325 x 00 . 7
V V P
dV P
PdV w
3
1 2
=
=
=
=
=

}
}
2
1
2
1
V
V
V
V


Example 2
Goates, and Ott, p. 12.
61
Calculate the work in J involved in the reversible compression of 10.0g of I He gas
from an initial pressure of 1.00atm to a final pressure of 5.00atm at the constant
temperature of 27C. Assume ideal gas behaviour.

Solution
) gas ideal , isothermal , reversible (
V
V
ln nRT
dV
V
nRT
w
V
nRT
P If
PdV w
2
1
V
V
V
V
2
1
2
1
=
=
=
=
}
}

Number of mol of He 50 . 2
00 . 4
00 . 10
= =

Ratio
2
1
1
2
1
2
P
P
P
nRT
P
nRT
V
V
= =

5.00
1.00
300ln x 8.314 x 2.50
P
P
nRTln
V
V
nRTln w
2
1
1
2
= = =

. J 10 x 00 . 1
4
=

5. Heat
Heat is energy that is exchanged between a system and its surroundings because of
differences in temperature only. Heat added to a system is given a positive sign, and
the normal symbol for heat is q.

Heat, like work, is not a state function. One of the significant early experiments in
this area was performed by Joule in the later 1840s. He showed that the temperature
of a water bath could be raised by the friction from a paddlewheel (doing work on
the system) as an alternative to placing the container over a flame (adding heat to the
system).

62

5.1 Heat capacity
The same amount of heat applied to different systems causes different changes in
temperature (a block of aluminium will increase more in temperature than a similar
block of ice for the same given amount of heat). This ability to change temperature
with applied heat is called heat capacity, defined as the amount of heat to raise the
temperature of the system by one degree Kelvin.

There are two forms of heat capacity.

Molar heat capacity
T n
q
C
A
=
(units of C: JK
-1
mol
-1
)

Specific heat capacity
T m
q
C
A
=
(units of C: JK
-1
g
-1
)

In the old calorie scale the heat capacity of water is 1 cal. K
-1
g
-1
. Thus the calorie is
the amount of heat needed to raise 1g of water 1C (or 1K).

(The above definitions of heat capacity are not precise - they assume heat capacity
does not vary with T. A more precise definition of C uses infinitesimal changes in q
and T.

63
i.e.
dT
dq
C =

However the above definitions will be sufficient for the cases we will be studying.)
Thus, providing there are no phase changes or chemical relations taking place,
measurements of temperature can be used to calculate q.

6. Summary
The forms of energy we are concerned with are

(i) thermal energy (heat)
(ii) mechanical energy (especially pressure-volume work)

In the following section we shall consider the relationships between heat and work.

7. The First Law of Thermodynamics
The First Law of Thermodynamics may be stated in a number of ways. e.g.

(i) The change in internal energy a of a system on going from an initial state A to a
final state B is determined only by the initial state A and the final state B and not
on the path by which the change was effected, or

(ii) All the energy which flows in and out of a system as heat or work must be
accounted for in the change in energy content of the system.

We will be most interested in the First Law quoted in the form of the following
equation:

U = q + w (11)

where U = change in internal energy
2



2
Some textbooks use E as the symbol for internal energy. U is the recommended symbol which is used in these notes. Choose either symbol and use
it consistently.
64
q = heat added to the system
w = work done on the system

Note that small letters, q and w, are used for changes which are dependent on
pathway, a capital letter U, is used for a property that depends only on the STATE
of the system. While q and w may each have many possible values, their sum (q + w)
is invariable and depends only on the initial and final states.

We often wish to express the First Law in terms of a very small (infinitesimal)
change. This statement is dU = dq + dw (12).

The First Law is really a restatement of the old idea of conservation of energy.
However, we have replaced the vague idea of energy with the precisely defined state
function, U. Thus the First Law is equivalent to the statement internal energy is a
state function.

Example
Barrow, G.M., Physical Chemistry (McGraw-Hill, 1973), p.121.
Calculate U for the conversion, at 100C and 1atm pressure, of 1 mole of water to
steam. The latent heat of vaporization of water is 40670J mol
-1
and the density of
liquid water can be taken as 1g mL
-1
and water vapour can be treated as an ideal gas.

Solution
Heat energy must be supplied to supply the latent heat of vaporization to convert the
liquid water to steam.

q = 40670J

Since PdV w
2
1
V
V
}
= (1)

and since P is constant

w = -P (V
2
- V
1
) (2)

Volume of 18g of water
65

= 18 x 10
-6
m
3


Volume of 1 mole of steam, at 100C and 101325 Nm
-2
is:

P
nRT
V =

101325
373 x 314 . 8 x 00 . 1
=

= 0.0306 m
3


Substitute in (2)

w = -101325 (0.0306 - 18 x 10
-6
)

= -3100J

Since U = q + w

U= 40670 - 3100

= 37570J mol
-1

8. Application of the First Law under special conditions

8.1 Processes occurring at constant volume
When a reaction is performed in a closed vessel, no volume changes can occur, so no
pressure-volume work may be done.

i.e. PdV w
2
1
V
V
}
=
= 0 at constant volume

From the First Law:
66

U = q + w

Since w = 0 at constant volume

U = q
v


where q
v
is heat absorbed by the system at constant volume.

8.2 Processes occurring at constant pressure
In general, the volume of the system involved in a chemical reaction in a vessel open
to the atmosphere changes as work is done on or by the surroundings. For a
macroscopic change the First Law is

U = q + w (11)

If work is restricted to pressure-volume work

PdV w
2
1
V
V
}
= (12)

At constant pressure

w = -P(V
2
- V
1
) (13)

(11) becomes

U = q
p
- P(V
2
- V
1
) (14)

where q
p
is heat absorbed at constant pressure. (14) may also be written

U
2
- U
1
= q
p
- P(V
2
- V
I
) (15)

U
2
= internal energy of the final state

67
U
1
= internal energy of initial state

Rearrange (15)

(U
2
+ PV
2
) - (U
1
+ PV
l
) = q
p
(l6)

At this stage it is convenient to introduce a new state function called ENTHALPY
(H) where:

H = U + PV (17)

It is easy to see that enthalpy is a state function because all of the species can the
right hand side of the defining equation (17) are state functions. In terms of enthalpy,
(16) becomes

H
2
- H
1
= q
p


or H = q
p
(18)

i.e. change in enthalpy = heat absorbed at constant pressure.

From an operational viewpoint, changes in properties are more significant than
absolute values. Since

H = U + PV (17)

H= U + (PV) (19)

At constant pressure, (19) becomes

H = U + PV

At constant volume, (19) becomes

H = U + VP

68
Worked example
G.M. Barrow, Physical Chemistry, (McGraw Hill, 1973) page 1, 140 Q.5
A chemical reaction in a gas mixture at 500C decreases the number of mole gas by
0.347. If the internal energy change is +23.8kJ, what is the enthalpy change? (Assume
the gases behave ideally).

Solution
If it is assumed that the reaction occurs at constant pressure, P, and that initially
there are n
1
mole of gas present:

Initially PV
1
= n
1
RT (1)

After reaction,

PV
2
= (n
1
- 0.347)RT (2)

Subtract (l) from (2)

P(V
2
- V
I
) = (n
1
- 0.347 - n
1
)RT

PV = - 0.347RT

By definition, enthalpy H is

H = U + PV

where U = internal energy.

If there is a change in enthalpy,

H = U + PV at constant pressure

i.e. H = 23.8 x 10
3
- 0.347 x 8.314 x 773.2

= 21.6kJ Ans.

69
8.3 Relationship between qP and qV
From the discussion above we know that:

U = q
v
at constant volume

and H = q
p
at constant pressure

but H = U + (PV) = U + PV at constant pressure

q
p
= q
v
+ PV (20)

i.e. the heat absorbed in a process taking place at constant pressure exceeds that
absorbed in a constant volume process by PV, since no mechanical work is done in
the constant volume process.

If the reactions under discussion involve only liquids and solids, the PV term is
usually negligible in comparison with the q terms. However, if gases are involved
PV is usually significant since changes in volume are often large.

Example
What is the difference between q
p
and q
v
for the following reaction carried out at
25C and one atmosphere pressure?

H
2
(g) + O
2
(g) H
2
0(l)

(Assume hydrogen and oxygen behave as ideal gases and that the density of water is
1.00g mL
-1
.)

Solution
q
p
= q
v
+ PV

For a reaction involving 1.00mol of H
2
, 0.500mol of O
2
and 1.00mol of water:

Volume of products = 18.0 x 10
-6
m
3

70

Volume of reactants =
101325
298 x 314 . 8 x 50 . 1


= 3.67 x 10
-2
m
3


Change in volume = V
2
- V
1

= 18.0 x 10
-6
- 3.67 x 10
-2


= - 3.67 x 10
-2
m
3

q
p
- q
v
= 101325 x -3.67 x 10
-2


= -3.72 x 10
3
Jmol
-1
Ans.

Where gaseous and condensed phase species both appear in a reaction it is generally
safe to consider only changes in volume of gaseous reactants.

Since PV = nRT

(PV) = (nRT)

At constant P, R and T

PV = RTn
q
p
- q
v
= RTn (21)

where n = number of mole of gaseous products

- number of mole of gaseous reactants

The problem studied above then becomes:

q
p
- q
v
= (0 - 1.5) x 8.314 x 298
71

= -3.72 x 10
3
Jmol
-1
Ans.

9. Heat capacities at constant pressure and constant volume
The heat capacity of a system is the amount of heat required to raise the temperature
of a system by one degree on the Kelvin scale.

i.e.
dT
dq
C = (22)

Where C = heat. capacity

dq = heat added to the system

dT = change in temperature

The heat capacity of a system is indefinite unless certain boundary conditions are
specified.

Thus at constant volume

v
v
dT
dq
C
|
.
|

\
|
= (23)

and at constant pressure,

P
p
dT
dq
C
|
.
|

\
|
= (24)

Since heat added to the system at constant volume is given by

dU = dq
v
for an infinitesimal change

72
then
V
v
dT
dU
C
|
.
|

\
|
= (25)

Since heat added to the system at constant pressure is given by

dH = dq
p
for an infinitesimal change

then
P
P
dT
dH
C
|
.
|

\
|
= (26)

For the special case of an ideal gas, internal energy is independent of volume and
enthalpy is independent of pressure, so that equations (25) and (26) become,

dT
dU
C
v
= (27)

dT
dH
C
p
= (28)

If C
p
and C
v
are independent of T, (27) and (28) become

U = nC
v
T (27a)

H = nC
P
T (28a)

10. Relationship between C
p
and C
v
for an ideal gas
From the First Law of Thermodynamics, for an infinitesimal change

dq = dU - dw (29)

But dw = -PdV for reversible pressure-volume work at constant pressure (30)

From (27)

dU = C
v
dT (31)
73

Substitute (30) and (31) into (29):

dq = C
v
dT + PdV (32)

For an ideal gas:

PV = RT for one mole

Differentiate with respect to T at constant pressure

R
dT
dV
P =

PdV = RdT (33)

Substitute (33) into (32)

RdT dT
v
C dq
p
+ = (34)

Divide (34) throughout by dT

R
v
C
dT
dq
p
+ = (35)

R
v
C
p
C + = (36)

where
p
C and
v
C are heat capacities for one mole of substance in each case.
Note: Adiabatic processes are not treated in this course. They are considered in
Physics 1. (Adiabatic processes are those where q = 0.)

74
11. Internal energy for an ideal gas
From the kinetic molecular theory the KE of an ideal gas is given by

KE =
2
3
RT (per mole) (37)

As this is the only energy possessed by an ideal gas

U = KE =
2
3
RT (38)

U =
2
3
RT

but U = C
v
T (31)

C
v
=
2
3
R (for an ideal gas)

C
P
= C
V
+ R C
P
=
2
5
R

These approximations can be used for real gases (unless the value is quoted in the
problem). Also, since U is a function only of T

T = 0 U = 0!

(for ideal gases only). It can also be shown that H = 0 also for T = 0. Joule
(1846)

75


The barrier is removed to allow the gas to expand into the evacuated half. If we
define the system to be the box enclosed by the insulated walls, then V = 0 (the
volume of the box doesn't change) w = 0.

Since the walls are insulated, q = 0, U = q + w = 0.

Since U = 0 T = 0.

Joule confirmed that there was no change in T (although for real gases there is a
small change due to the attractive forces between molecules; for ideal gases these are
assumed to be zero. Joule's apparatus was not sensitive enough to measure these
changes.

Example
Consider the cyclic change carried out on one mole of an ideal gas

Calculate U, H, q and w for each pathway, and for the whole cycle.

76
Step I
This is an isochoric (constant volume) step

v = 0

w = 0

q = n C
V
T = C
V
T (n = 1)

=
2
3
R(600 - 300) = 3741J

U = q + W = q
V
= 3741J

H = U + (PV) = U + VP (since V is constant) = 3741 + .02494

(200,000 - 100,000) = 6235J

Step 2

P = 0 (isobaric)

q = n C
P
T =
2
5
R (300 - 600)

= -6235J (watch sign convention!)

= H (since H = q
P
)

w = - PV = -200,000 (.01247 - .02494)

= 2494J

U = q + w = 2494 - 6235 = -3741J

Step 3
T = 0 (isothermal)

U = 0 = H
77


12.47
24.94
ln 300 x 8.314
V
V
ln nRT w
1
2
= =

= -1729J

Since q + w = U = 0

q = 1729J

for the whole cycle

U = 3741 - 3741 + 0 = 0

H = 6235 - 6235 = 0

w = 0 + 2494 - 1729 = 765 J

q = 3741 - 6235 + 1729 = -765 J

confirms U and H are state functions but q and w are not (since this is a cyclic
process).

The formulae can be summarised in the table below (the formulae for the reversible,
isothermal case apply to ideal gases only - the other two cases are more general).




V= 0 P= 0 T= 0
q nC
V
T nC
P
T nRT ln
1
2
V
V

w 0 -PV -nRT ln
1
2
V
V

U nC
V
T nC
P
T - PV 0

78
H nC
V
T + VP nC
P
T 0

Вам также может понравиться