You are on page 1of 28

LECTURE 9 THE SECOND LAW OF THERMODYNAMICS

9.1. FIRST VS. SECOND LAW

First Law of Thermodynamics:


The energy of the universe is constant - energy is conserved.
 This law tells us nothing about the spontaneity of physical and chemical transformations.

Consider the formation of water:


H2(g) + ½O2(g) → H2O(l) ΔfHo = -286 kJ mol-1

Using the first law, it is easy to calculate ΔU and ΔH associated with this spontaneous reaction - we
can also calculate ΔU and ΔH for the reverse reaction, which we know does not occur spontaneously.

What about gases?


We know that gases expand spontaneously to fill a container - we know that the opposite does not
happen - unless some sort of work is done to bring about this reverse change.
Neither the sign nor the magnitude of ΔU and ΔH tell us which way the reaction will go - however, we
can get this information from the Second Law of Thermodynamics.
The first law gives us no clue what processes will actually occur and which ones will not. For that
matter, why does anything ever happen at all? The universe is an isolated system after all. There is no change
in internal energy. There is no heat transferred in or out and no work is done on or by the system as a whole.
q=0; w=0; ΔU=0.

Without only the first law, and without the second law:
The universe is a boring place.

Yet, the universe is not a boring place:


 Stars come into existence and blow up into oblivion.
 Planets are created and hurl around stars.
 Life evolves amid all of this turmoil on these planets.

Why? What is driving all of this?


These processes do not lead to a lower energy for the universe as a whole. So what is going on?

9.2. STATEMENTS OF THE SECOND LAW

Elements in a closed system tend to seek their most probable distribution; in a closed system entropy
always increases.
1. Clausius (1822-1888) - It is impossible that, at the end of a cycle of changes, heat has been
transferred from a colder to a hotter body without at the same time converting a certain amount of
work into heat.
2. Lord Kelvin (1824-1907) - In a cycle of processes, it is impossible to transfer heat from a heat
reservoir and convert it all into work, without at the same time transferring a certain amount of heat
from a hotter to a colder body.
3. Ludwig Boltzmann (1844-1906) - For an adiabatically enclosed system, the entropy can never
decrease. Therefore, a high level of organization is very improbable.
4. Max Planck (1858-1947) - A perpetual motion machine of the second kind is impossible.
5. Caratheodory (1885-1955) - Arbitrarily near to any given state there exist states which cannot be
reached by means of adiabatic processes.

RVM Lecture Manual: Physical Chemistry 1 1


9.3. THE SECOND LAW OF THERMODYNAMICS

The Second Law of Thermodynamics can be stated in a number of equivalent ways; an early
statement of the law from Lord Kelvin says: No process is possible in which the sole result is the absorption of
heat from a reservoir and its complete conversion into work .

All heat engines have a source of heat and a cold sink,


into which some heat is always discarded and not converted into
work.

The engine on the left violates the second law - this is


equivalent to a ball initially at rest on a heated surface starting to
bounce (i.e., heat absorbed from this surface is converted into
orderly work) - this has not been observed.

What determines how physical and chemical processes


proceed spontaneously? Energy? Tendency towards minimum
energy? It is none of these, as we shall see.

SIMPLE PROCESSES

Process 1: First Law fulfilled


 A perfectly elastic ball in a vacuum is dropped from
some height with an initial potential energy.
 When it strikes the ground, all energy is converted into
kinetic energy.
 The ball bounces back to the original height, where all
of the kinetic energy is converted back to the original
potential energy.

Process 2: Why doesn’t the egg bounce?


 An egg is dropped from the same height as the ball.
 Initial and final states are not the same - seem to have
lost energy!
 First Law: we cannot lose energy - where did it go?
- Converted into random molecular motion and heat
- We wound up with a more disorganized form of matter

The universe tends to more random disorganized states.

DISPERSAL OF ENERGY

Spontaneous changes are accompanied by dispersal of energy


into a more disordered form.

Bouncing ball:
 Ball does not return to original height, as energy is distributed
to molecules in the floor and ball as heat - inelastic energy
losses.
 Ball eventually comes to rest, losing all energy into thermal
motion of atoms in the floor - the reverse process will never occur!

RVM Lecture Manual: Physical Chemistry 1 2


Heated floor and ball:
 Molecules and atoms in the floor (and ball) undergo thermal
(random) motion.
 In order for the ball to spontaneously bounce, it would require a
spontaneous localization of motion, to create an upwards motion of all of
the atoms - a virtually impossible process.

9.4. IRONY OF THE SECOND LAW: ENTROPY

The universe tends towards more random, disorganized states.

This is a rather loose statement of the second law of thermodynamics and our way of quantitating the
disorder and randomized motion in one state versus another is a state function called entropy:

Increasing entropy means increasing disorder and randomized motion.

Melodramatic viewpoint:
Every star that burns, every planet whose orbit is slowly decaying, every breath you take and calorie
you metabolize brings the universe closer and closer to the point when the entropy is maximized, organized
movement of any kind ceases, and nothing ever happens again. There is no escape. No matter how
magnificent life in the universe becomes or how advanced, the slow increase in entropy cannot be stopped -
the universe will eventually die.

Irony of the second law:


Without the second law, nothing would ever happen at all. With it, the universe is doomed.

9.5. GENERAL THOUGHTS ON SECOND LAW

Signpost for spontaneous change:


Direction of change that leads to general chaotic dispersal of total energy of an isolated system.
 Ball will bounce and come to rest.
 Gas spontaneously expands, does not spontaneously contract.
 Object does not suddenly become warmer than its surroundings for no reason at all -
random vibration of atoms will not suddenly lead to an excess of thermal motion (and
therefore heating) - however, the object does spontaneously release energy into the
surroundings as thermal motion (consequence of chaos).

Puzzling thought: The fall into disorder can result in highly order substances such as crystals, proteins, life!
etc.: organized structures and patterns can emerge as energy and matter disperse (i.e., the entropy increases) .

The meaning of life: to increase entropy in the universe.

9.6. SPONTANEITY AND REVERSIBILITY

Spontaneous: It has little to do with an impromptu gesture, a witty saying or impulsive shopping. It means a
process that results in a change from one state to another in an irreversible way. Anything that happens in the
universe that results in an irreversible change in state is spontaneous .

RVM Lecture Manual: Physical Chemistry 1 3


Reversible Change (Not Spontaneous):
Truly reversible processes do not happen in reality, because in a truly reversible process all forces
would be perfectly balanced and there would be no driving force for the system to move. By moving things
very slowly always keeping forces in near perfect balance, we can approximate reversible processes to
whatever degree we like. For example, during reversible expansion of a gas, we keep the pressures essentially
the same on the inside and the outside - if this was strictly true, the gas would not have any driving force to
expand and nothing would ever happen. However, we can make it as close to true as we like by making the
imbalance as small as we want.

Irreversible Change (Spontaneous):


All processes that really happen are irreversible; forces driving process are substantially out of
balance.

9.7. SECOND LAW AND ENTROPY

The law used to identify spontaneous change, and can be quantified in terms of a state function
known as entropy, S.

Second Law
First Law Uses entropy, S, to identify the
Uses internal energy, U, to spontaneous (irreversible)
identify permissible changes changes
among the permissible changes

The entropy of an isolated system increases in the course of spontaneous change.


ΔStotal > 0
where Stot is the total entropy of a system and its surroundings.
i.e., ΔStotal = ΔSsys + ΔSsur

Entropy is a measure of molecular disorder in a system, letting us assess whether one state is
accessible from another via spontaneous change.

9.8. THERMODYNAMIC DEFINITION OF ENTROPY

ENTROPY
“Heat cannot spontaneously flow from a colder location to a hotter location.”
There are many versions of the second law, but they all have the same effect, which is to explain the
phenomenon of irreversibility in nature.
There are several ways in which the second law of thermodynamics can be stated. Listed below are
three that are often encountered. As describes in class (and as derived in almost every thermodynamics
textbook), although the three may not appear to have much connection with each other, they are equivalent.

1. “No process is possible whose sole result is the absorption of heat from a reservoir and the
conversion of this heat into work.” (Kelvin-Planck statement of the second law)
T2
System Q
W

T1

RVM Lecture Manual: Physical Chemistry 1 4


2. “No process is possible whose sole result is the transfer of heat from a cooler to a hotter body.”
(Clausius statement of the second law)

T2 Q

T1 < T2 Not possible


T1
Q

3. There exists for every system in equilibrium a property called entropy, S, which is a thermodynamic
property of a system. For a reversible process, changes in this property are given by
𝑑𝑞𝑟
𝑑𝑆 =
𝑇
The entropy change of any system and its surroundings, considered together, is positive and
approaches zero for any process which approaches reversibility.

ΔStotal ≥ 0
For an isolated system, i.e., a system that has no interaction with the surroundings, changes in
the system have no effect on the surroundings. In this case, we need to consider the system only, and the
first and second laws become:
ΔEsystem = 0
ΔSsystem ≥ 0
All of these statements are equivalent, but 3 give a direct, quantitative measure of the departure
from reversibility.

An infinitesimal change in entropy is dS, and this can occur as a result of a chemical or physical
process. The thermodynamic definition of entropy:

𝑓
𝑑𝑞𝑟𝑒𝑣 𝑑𝑞𝑟𝑒𝑣
𝑑𝑆 = ∆S = ∫
𝑇 𝑇
𝑖
Infinitesimal change measurable change

If we perform some process along a reversible path, then the entropy produced should be
proportional to the amount of heat produced (or consumed) and inversely proportional to the
temperature.

Measurable entropy difference between two states , ΔS: find a reversible path, calculate heat
supplied at each stage of the path divided by temperature at which heat is supplied.

Why is this? The change in the extent to which energy is dispersed in a random disorderly
manner depends on the amount of energy transferred as heat (creates random “thermal” motion). Work
is not accompanied by increase in random motion, but rather, implies uniform motion, and is therefore
not accompanied by changes in entropy.

RVM Lecture Manual: Physical Chemistry 1 5


ENTROPY EXAMPLE: ISOTHERMAL EXPANSION

Consider reversible isothermal expansion of a perfect gas (pV = nRT).

p = pex ΔT = 0 ΔV = Vf - Vi , w < 0

In an isothermal expansion, ΔU = 0, so q = -w.

Since T is constant:
𝑓
1 𝑞𝑟𝑒𝑣
∆𝑆 = ∫ 𝑑𝑞𝑟𝑒𝑣 =
𝑇 𝑇
𝑖

We know the work of this type of expansion, so:


𝑉𝑓
𝑞𝑟𝑒𝑣 = −𝑤𝑟𝑒𝑣 = 𝑛𝑅𝑇𝑙𝑛 ( )
𝑉𝑖
Thus, it follows that,
𝑉𝑓
∆𝑆 = 𝑛𝑅𝑙𝑛 ( )
𝑉𝑖

Simple probability: molecules will randomly occupy the space that is available to them - tendency of
system to explore all available states.

9.9. SURROUNDINGS

Surroundings are treated in a manner similar to the system, except that surroundings are so large
that we consider them to be isothermal and (generally) at constant pressure. Thus, qsur = ΔHsur (recall, ΔH =
qp).
Therefore, heat transferred to the surroundings is equal to a state function and is independent of the
path that the heat used to get out into the surroundings: it has been transferred by a reversible path.
Another view: the surroundings are generally assumed not to change state when something happens
in the system, thus transfer of heat to and from the surroundings is effectively reversible (since there is no
change of state, we cannot say that the change was irreversible):
𝑑𝑞𝑠𝑢𝑟,𝑟𝑒𝑣 𝑑𝑞𝑠𝑢𝑟
𝑑𝑆𝑠𝑢𝑟 = =
𝑇𝑠𝑢𝑟 𝑇𝑠𝑢𝑟
Since the surroundings are isothermal:
𝑞𝑠𝑢𝑟
∆𝑆 =
𝑇𝑠𝑢𝑟
So for any adiabatic change (qsur = 0):
ΔSsur = 0

9.10. ENTROPY AS A STATE FUNCTION

In order to prove that entropy is a state function, the integral of


dS must be independent of path:
𝑑𝑞𝑟𝑒𝑣
∮ =0
𝑇
where ∮ means integration about a closed path.

If the integration of the above equation above an arbitrary cycle


(cyclic system) is zero, this will show that the entropy of the system in
the initial and final states is the same, regardless of the path taken.
In a thermodynamic cycle, the overall change in a state function is zero, and is independent of path.

RVM Lecture Manual: Physical Chemistry 1 6


A. Change in Entropy of a Heat Reservoir
Heat reservoir is a body of infinite heat capacity. The change in entropy of a heat reservoir is
computed using
𝑄
∆𝑆 =
𝑇

B. Entropy Balance for a Control Volume


Time rate entropy change = net rate of energy transfer due to mass flow – net rate of entropy
transfer due to heat transfer + rate of entropy production due to irreversibilities.
𝑛
𝑑𝑆𝑐𝑣 𝑄𝑗
= ∑ 𝑚𝑖 𝑆𝑖 − ∑ 𝑚𝑜 𝑆𝑜 + ∑ + 𝜎𝑐𝑣
𝑑𝑡 𝑇𝑗
𝑖𝑛 𝑜𝑢𝑡 𝑗=1

C. Change in Entropy of an Ideal Gas


The entropy of the system depends only on the initial and final states of the system, and hence, as
in the case of E and H we may write that the change in entropy of a process is:
ΔS = S2 – S1
Where S2 and S1 are, respectively, the entropies of the system in the final and initial states.
Furthermore, the differential change in S, dS, is given by:
𝑑𝑞𝑟
𝑑𝑆 =
𝑇
Where dqr is the infinitesimal quantity of heat absorbed in a process taking place under
reversible conditions at temperature T. In the case of a finite reversible change at constant temperature,
dS becomes ΔS, dqr, becomes qr, and eq.(2) takes the form:
𝑞𝑟
∆𝑆 =
𝑇
Process Governing Equation
𝑇2
𝐶𝑣 𝑑𝑇 𝑉2
∆𝑆 = ∫ + 𝑅𝑙𝑛
𝑇1 𝑇 𝑉1
Any process 𝑇2 𝐶 𝑑𝑇
𝑝 𝑃2
∆𝑆 = ∫ + 𝑅𝑙𝑛
𝑇1 𝑇 𝑃1
𝑉 𝑃
Isothermal expansion into a vacuum ∆𝑆 = 𝑛R ln( 2) = nR ln( 2)
𝑉1 𝑃1

𝑇2
𝑑𝑇
Isobaric ∆𝑆 = ∫ 𝐶𝑝
𝑇1 𝑇
𝑇2
𝑑𝑇
Isochoric ∆𝑆 = ∫ 𝐶𝑣
𝑇1 𝑇

Isentropic
 Process is reversible and adiabatic
dQrev = 0, therefore, dSt = 0.
Entropy is constant during a reversible adiabatic process.

Entropy of Cycle
ΔScycle = 0, reversible
ΔScycle > 0, irreversible
ΔScycle <0, impossible

Irreversible, treat them as reversible and perform ΔS t =∫(dQrev/T), then apply relationship of reversible to
irreversible.

RVM Lecture Manual: Physical Chemistry 1 7


Example:
Calculate the standard entropies for the following reactions at 25°C:
a. N2 (g) + 3H2 (g) →2NH3 (g)
b. N2O4 (g) → 2NO2 (g)

Solution:
We will first write the balanced equations for each reaction. Then apply Eq. 3.69 in order to
determine the standard entropy changes for both A and B.
a. (1) N2 (g) + 3H2 (g) →2NH3 (g)
ΔS° = ∑ΔS°(products) - ΔS°(reactants)
ΔS° = 2ΔS°(NH3, g) – (3ΔS°(H2, g) + ΔS°(N2, g))
ΔS° = 2ΔS°(NH3, g) - 3ΔS°(H2, g) - ΔS°(N2, g)
ΔS° = [(2 x 192.45) – (3x 130.68) – (191.61)]JK-1
ΔS° = -198.75 JK-1mol-1
b. (2) N2O4 (g) → 2NO2 (g)
ΔS° = ∑ ΔS°(products) - ΔS°(reactants)
ΔS° = 2ΔS°(NO2,g) - ΔS°(N2O4,g)
ΔS° = [(2 x 240.1) – 304.2] JK-1
ΔS° = 176 JK-1mol-1
_________________________________________________________________________________________________________________________

9.11. CARNOT CYCLE


The Carnot cycle, named after Sadi Carnot, has four
reversible stages.
1. Reversible isothermal expansion from A to B at Th, ΔS =
qh/Th, qh is heat supplied from a hot source, and is
positive.
2. Reversible adiabatic expansion from B to C, no heat
leaves system (ΔS = 0), temperature falls from Th to Tc,
where Tc is the temperature of the cold sink.
3. Reversible isothermal compression from C to D, ΔS =
qc/Tc, qc is heat released into a cold sink, and is
negative.
4. Reversible adiabatic compression from D to A, no heat
enters the system (ΔS = 0), temperature rises from Tc
to Th.

Total entropy change in Carnot cycle:


𝑞ℎ 𝑞𝑐
∮ 𝑑𝑆 = +
𝑇ℎ 𝑇𝑐
This is zero since:
𝑞ℎ 𝑇ℎ
=−
𝑞𝑐 𝑇𝑐

Carnot Cycle
1. The thermal efficiency of an internally irreversible heat engine is always less than the thermal
efficiency of an internally reversible heat engine operating with heat transfer at the same high-
and low-temperature regions.
2. The thermal efficiencies of two internally reversible heat engines operating with heat transfer at
the same high- and low-temperature regions are equal.

RVM Lecture Manual: Physical Chemistry 1 8


ΔS=0 FOR CARNOT CYCLE
Recall for reversible isothermal expansion of a perfect gas:
𝑉𝐵 𝑉𝐷
𝑞ℎ = 𝑛𝑅𝑇ℎ 𝑙𝑛 ( ) 𝑞𝑐 = 𝑛𝑅𝑇𝑐 𝑙𝑛 ( )
𝑉𝐴 𝑉𝐶
and for reversible adiabatic processes:
VAThc = VDTcc VCTcc = VBThc

Recall:
𝐶𝑣,𝑚
VfTf c = ViTic 𝑐=
𝑅
Multiplying the expressions together:
VAVCThcTcc = VDVBThcTcc
and simplifying:
𝑉𝐴 𝑉𝐷
=
𝑉𝐵 𝑉𝐶
we get:
𝑉𝐴
𝑞𝑐 = 𝑛𝑅𝑇𝑐 𝑙𝑛 ( )
𝑉𝐵
Therefore:
𝑞ℎ 𝑇ℎ
=−
𝑞𝑐 𝑇𝑐

9.12. EFFICIENCY
The same calculation just completed for gases applies to all
types of materials and systems. We define the efficiency, ε, of a heat
engine:
work performed ‖w‖
ε= =
heat abosrbed qh
The greater the work output from a given supply of heat, the
greater the efficiency of the engine.
Work performed by the engine is the difference between heat
supplied from the heat source and returned to the cold sink:
𝑞ℎ + 𝑞𝑐 𝑞𝑐
𝜀= = 1+
𝑞ℎ 𝑞ℎ
Since qc < 0,
𝑇𝑐
𝜀𝑟𝑒𝑣 = 1 −
𝑇ℎ

CARNOT CYCLES AND ENGINES


Second Law: all reversible engines have the same
efficiency regardless of construction - Two engines A and B,
assume A more efficient than B, coupled together using the same
reservoirs.
A: takes heat qh, releases heat qc, B: takes heat qc
releases qh Since A more efficient than B, not all work A produces
is needed for this process, and difference can be used to do work.
Net result: cold reservoir stays the same, hot reservoir
loses d[energy, and work has been produced.
Problem: heat (disordered thermal motion) has been
converted directly into work (ordered molecular motion)
without the need for a cold sink - assumption that reversible
engines A and B can have different efficiencies is false!
Relationship between heat transfers and temperatures
must be independent of working materials.

RVM Lecture Manual: Physical Chemistry 1 9


CARNOT CYCLES: S IS A STATE FUNCTION

Any reversible cycle can be thought of as a collection


of Carnot cycles - this approximation becomes exact as cycles
become infinitesimal.
Entropy change around an individual cycle is zero.
Sum of entropy changes over all cycles is zero.
In interior, entropy change along any path is
cancelled by the entropy change along the path of shared with
its neighbor.
Thus, all entropy changes cancel except those along
the perimeter of the entire cycle - but the sum of the paths
along of the perimeter, if the paths are infinitesimal in size,
match the overall cycle exactly, so,
𝑞𝑟𝑒𝑣 𝑞𝑟𝑒𝑣
∑ = ∑ =0
𝑇 𝑇
𝑎𝑙𝑙 𝑝𝑒𝑟𝑖𝑚𝑒𝑡𝑒𝑟
𝑑𝑞𝑟𝑒𝑣
∮ =0
𝑇
dS is an exact differential and therefore S is a state function.

RUDOLF CLAUSIUS
Rudolf Julius Emanuel Clausius (January 2, 1822 – August 24, 1888),
was a German physicist and mathematician, and was one of the founders of
thermodynamics. His most important paper, on the mechanical theory of
heat, published in 1850, first stated the basic ideas of the second law of
thermodynamics. In 1865 he introduced the concept of entropy. He
discovered the fact that entropy can never decrease in a physical process and
can only remain constant in a reversible process, a result which became
known as the Second Law of Thermodynamics.
Clausius graduated from the University of Berlin in 1844, and got his
doctorate from the University of Halle in 1848. He then taught in Berlin,
Zürich, Würzburg, and Bonn.
In 1870 Clausius organized an ambulance corps in the Franco-
Prussian War. He was wounded in battle, leaving him with a lasting disability. He was awarded the Iron Cross
for his services. His wife, Adelheid Rimpham, died in childbirth in 1875, leaving him to raise their six children.
He continued to teach, but had less time for research thereafter. A crater on the Moon has been named in
honor of this founding father of thermodynamics.

9.13. THE CLAUSIUS INEQUALITY


Isothermal expansion of a perfect gas:

𝑉𝑓
𝑤isothermal,rev = 𝑛𝑅𝑇𝑙𝑛 ( )
𝑉𝑖
𝑤isothermal,irrev = −𝑝f ∆𝑉
𝑛𝑅𝑇
=− ∆𝑉
𝑉𝑓
∆𝑉
= −𝑛𝑅𝑇
𝑉𝑓

RVM Lecture Manual: Physical Chemistry 1 10


 Take Vf = 2Vi, wirrev = -(0.5)nRT, and wrev = -(0.693)nRT (more work done on surroundings by
reversible expansion)
 Since isothermal expansion, ΔU = ΔH, and q = -w
 This means that qrev > qirrev
 Ssys is state function, independent of path, ΔSsys = qrev/T = (0.693)nR
 Ssur depends on amount of heat transferred: ΔSsur,rev = -(0.693)nR, ΔSsur,irrev = -(0.5)nR

 In the reversible case, ΔStot,rev = 0, since the entropy change of system and surroundings will have
opposite signs.
 In the irreversible case, ΔStot,irrev = 0.193nR

This is a general result: all irreversible reactions are spontaneous and have a total entropy change
which is greater than zero.
For system in mechanical and thermal contact with surroundings, at some temperature, T, there may
not be mechanical equilibrium. Any change in state has entropy changes dSsys(system) & dSsur(surroundings).
The process may be irreversible, total entropy will increase.
𝑑𝑆𝑠𝑦𝑠 + 𝑑𝑆𝑠𝑢𝑟 ≥ 0, 𝑜𝑟 𝑑𝑆𝑠𝑦𝑠 ≥ −𝑑𝑆𝑠𝑢𝑟
Since ΔSsur = qsur/Tsur, dSsur = -dq/T, where dq is the heat provided to the system during the process,
and dqsur = -dq). Then for any change,
𝑑𝑞
𝑑𝑆𝑠𝑦𝑠 ≥
𝑇

dSsys ≥ 0 (system isolated from surroundings)

This is our signpost of spontaneous change: in an isolated system the entropy of the system alone
cannot decrease due to spontaneous change.

Example 1: Irreversible adiabatic change, dq = 0, and therefore,


dS ≥ 0
Entropy of system increases, entropy of surroundings stays
constant (since no heat transfer has occurred), dStot ≥ 0.

Example 2: Irreversible isothermal expansion of perfect gas: dq


= -dw
If into a vacuum, w = q = 0, so dS ≥ 0, dSsur = 0 and dStot ≥ 0.

Example 3: Spontaneous cooling

|𝑑𝑞| |𝑑𝑞| 1 1
𝑑𝑆 = − = |𝑑𝑞| ( − )
𝑇𝑐 𝑇ℎ 𝑇𝑐 𝑇ℎ

dS > 0, since Th > Tc, so cooling is spontaneous - if two


temperatures are equal, dStot = 0.

9.14. ENERGY QUANTIZATION

At the start of this century, there were certain physical anomalies which could not be explained using
Newtonian (Classical) Mechanics - at the atomic level matter behaves differently.
A new set of mechanics, developed by Einstein, Schroedinger and others, demonstrated both
experimentally and theoretically that at the atomic and molecular levels, the energy of particles is not a
continuum, but rather, is quantized: Quantum Mechanics.

RVM Lecture Manual: Physical Chemistry 1 11


Translational motion of molecules in macroscopic containers can be treated in many cases by
classical mechanics, but rotation, vibration and motion of electrons (electronic transitions) have quantized
energy levels.

9.15. POPULATIONS OF STATES

Molecules and atoms at a single temperature will individually be in one energy state in one instant,
and another completely different energy state in another, due to collisions, etc. We cannot track the energy of
an individual molecule: but, we can however monitor the populations of the various energy states quite easily
(i.e., watch the behaviour of collections or ensembles of molecules).
At temperatures > 0, molecules are distributed over available energy levels according to the
Boltzmann Distribution, which gives the ratio of particles in each energy state:
𝑁𝑖
= 𝑒 −(𝐸𝑖−𝐸𝑗 )/𝑘𝑇
𝑁𝑗
At the lowest temperature T = 0, only the lowest energy state is occupied. At infinite temperature, all
states are equally occupied.
Degenerate states: States which have the same energy. These will be equally populated!

9.16. BOLTZMANN DISTRIBUTIONS

Populations for (a) low & (b) high


temperatures Boltzmann predicts an
exponential decrease in population with
increasing temperature

At room T, only the ground electronic state is


populated. However, many rotational states
are populated, since the energy levels are so
closely spaced.

More states are significantly populated if


energy level spacings are near kT!

RVM Lecture Manual: Physical Chemistry 1 12


9.17. BOLTZMANN’S ENTROPY

Ludwig Boltzmann (1844-1906) was an Austrian mathematician


and physicist who was most famous for his creation of statistical mechanics,
which connects the properties and behaviour of collections or ensembles of
atoms and molecules with the large scale properties and behaviour of the
substances of which they were the building blocks.

Entropy is a measure of disorder in a system, where disorder is


defined formally as the number of different microstates a system can be in,
given that the system has fixed composition, volume, energy, pressure and
temperature. By "microscopic states", we mean the exact states of each of
the molecules making up the system.

S = klnW or W = eS/k

where W is defined as the number of total possible microstates or


the number of total ways the constituents of the system can combine. Thus,
the more microstates or combination pathways there are, the higher the entropy. Boltzmann has also be
called a “grandfather” to quantum theory, since modern stat mech connects classical mechanics and quantum
mechanics.

9.18. MAXWELL’S DEMON

James Clerk Maxwell (1831-1879) was a Scottish mathematician


and physicist who published physical and mathematical theories of the
electromagnetic field, published as “A Treatise on Electricity and Magnetism
(1873)”, which included the formulas today known as the Maxwell
equations.

With Clausius, he developed the kinetic theory of gases. In


"Illustrations of the Dynamical Theory of Gases" (1860), he described the
velocity distribution of molecules. His studies of kinetic theory led him to
propose the Maxwell's demon paradox in an 1867 letter to Tait.

Maxwell's demon (termed a "finite being" by Maxwell) is a


tiny hypothetical creature that can see individual molecules. The
demon can make heat flow from a cold body to a hot one by opening a
door whenever a molecule with above average kinetic energy
approaches from the cold body, or below average kinetic energy
approaches from the hot body, then quickly closing it. This process
appears to violate the second law of thermodynamics, but was used by
Maxwell to show that the second law of thermodynamics is a
statistical law describing the properties of a large number of particles.

Why?
Because the acts of observation and information necessarily requires changes in entropy
(information theory - entropy is the lack of information about a system!).

RVM Lecture Manual: Physical Chemistry 1 13


EXERCISES:
1. Calculate the change in entropy when 50 k] of energy is transferred reversibly and isothermally as heat to
a large block of copper at (a) 0°C, (b) 70°C.
2. Calculate the molar entropy of a constant-volume sample of argon at 250 K given that it is 154.84 J K-1
mol-1 at 298 K.
3. Calculate ΔS (for the system) when the state of 2.00 mol diatomic perfect gas molecules, for which Cp,m
= 7/2R, is changed from 25°C and 1.50 atm to 135°C and 7.00 atm. How do you rationalize the sign of ΔS?
4. A sample consisting of 2.00 mol of diatomic perfect gas molecules at 250 K is compressed reversibly and
adiabatically until its temperature reaches 300 K. Given that Cv,m = 27.5 J K-1mol-1, calculate q, w, ΔU, ΔH,
and ΔS.
5. Calculate ΔH and ΔStot when two iron blocks, each of mass 1.00 kg, one at 200°C and the other at 25°C, are
placed in contact in an isolated container. The specific heat capacity of iron is 0.449 J K-1 g-1 and may be
assumed constant over the temperature range involved.
6. Consider a system consisting of 1.5 mol CO2 (g), initially at 15°C and 9.0 atm and confined to a cylinder of
cross-section 100.0 cm2. The sample is allowed to expand adiabatically against an external pressure of 1.5
atm until the piston has moved outwards through 15 cm. Assume that carbon dioxidemay be considered
a perfect gas with CV,m = 28.8 J K-l mol-1, and calculate (a) q, (b) w, (c) ΔU, (d) ΔT, (e) ΔS.
7. The enthalpy of vaporization of methanol is 35.27 k] mol-I at its normal boiling point of 64.1 "C. Calculate
(a) the entropy of vaporization of methanol at this temperature and (b) the entropy change of the
surroundings.
8. Calculate the standard reaction entropy at 298 K of:
(a) Zn(s) + Cu2+ (aq) → Zn2+ (aq) + Cu(s) (b)C12 H22 O11 (s) + 12O2 (g) + 12CO2 (g) + 11H2 O(l)
9. Combine the reaction entropies calculated in Exercise 8 with the reaction enthalpies, and calculate the
standard reaction Gibbs energies at 298 K.
10. Use standard Gibbs energies of formation to calculate the standard reaction Gibbs energies at 298 K of
the reactions in Exercise 8.
11. Calculate the standard Gibbs energy of the reaction CO(g) + CH 3OH(l)→CH3COOH(l) at 298 K, from the
standard entropies and enthalpies of formation given in the Data section.
12. The standard enthalpy of combustion of solid urea (CO(NH2)2) is −632 kJ mol−1 at 298 K and its
standard molar entropy is 104.60 J K−1 mol−1. Calculate the standard Gibbs energy of formation of urea
at 298 K.
13. Calculate the change in the entropies of the system and the surroundings, and the total change in entropy,
when the volume of a sample of argon gas of mass 21 g at 298 K and 1.50 bar increases from 1.20 dm3 to
4.60 dm3 in (a) an isothermal reversible expansion, (b) an isothermal irreversible expansion against pex
= 0, and (c) an adiabatic reversible expansion.
14. Calculate the maximum non-expansion work per mole that may be obtained from a fuel cell in which the
chemical reaction is the combustion of propane at 298 K.
15. A certain heat engine operates between 1000 K and 500 K. (a) What is the maximum efficiency of the
engine? (b) Calculate the maximum work that can be done by for each 1.0 kJ of heat supplied by the hot
source. (c) How much heat is discharged into the cold sink in a reversible process for each 1.0 kJ supplied
by the hot source?

ENTROPY CHANGES AND PROCESSES

9.19. ENTROPY AT A PHASE TRANSITION

Changes in molecular order occur when a substance freezes or boils.


Consider the phase transitions of water, at transition temperatures Ttrs. For ice water, Ttrs = 273 K,
ice in equilibrium with liquid water at 1 atm and boiling water, Ttrs = 373 K, liquid water in equilbrium with
vapour at 1 atm.

RVM Lecture Manual: Physical Chemistry 1 14


The external pressure is constant for a glass of ice water, and in order
to match attractive forces between ice molecules, energy must come from
kinetic energy of the water molecules or the surroundings.
At Ttrs, any transfer of heat between the system and surroundings is
reversible since the two phases in the system are in equilibrium (the forces
pushing the ice towards melting are equal to those pushing the water towards
freezing) - so a phase transition is reversible.
It does not matter how the ice melts (what path it takes) since
entropy is a state function. What does matter for this particular expression is
that the system be isothermal. If it was not isothermal, one would have a
problem examining the process in steps – as we shall see.

At constant pressure, q = ΔtrsH, and the change in molar entropy is


∆𝑡𝑟𝑠 𝐻
∆𝑡𝑟𝑠 𝑆 =
𝑇𝑡𝑟𝑠
For exothermic and endothermic phase transitions:

Consistent with decreasing disorder: gases > liquids > solids

Example: when compact condense phase vapourizes into a widely dispersed gas, one can expect
an increase in the disorder of the molecules.

9.20. TROUTON’S RULE

Trouton’s Rule: This empirical observation states that most liquids have approximately the same
standard entropy of vaporization, ΔvapSo. 85 J K-1 mol-1: So, ΔvapHo = Tb × 85 J K-1 mol-1.

Some Standard Entropies of Vaporization


ΔvapHo(kJ mol-1) θboil(oC) ΔvapSo(J K-1 mol-1)
Benzene +30.8 80.1 +87.2
CCl4 +30.00 76.7 +85.8
Cyclohexane +30.1 80.7 +85.1
H2S +18.7 -60.4 +87.9
Methane +8.18 -161.5 +73.2
Water +40.7 100.0 +109.1

Exceptions:
 In water, molecules are more organized in the liquid phase (due to hydrogen bonding), so a greater
change of disorder occurs upon vapourization.
 In methane, the entropy of the gas is slightly low (186 J K-1 mol-1 at 298 K) and in light molecules very few
rotational states are accessible at room temperature - associated disorder is low.

9.21. EXPANSION OF A PERFECT GAS

The ΔS for an isothermally expanding perfect gas can be written as

RVM Lecture Manual: Physical Chemistry 1 15


𝑉𝑓
∆𝑆 = 𝑛𝑅𝑙𝑛 ( )
𝑉𝑖

(equation applies for reversible or irreversible change, S = state function)


N.B.: Atkins uses ΔS instead of ΔSsys - they are equivalent!

Reversible change:
ΔStot = 0
The surroundings are in thermal and mechanical equilibrium with system, so Δ Ssur = -ΔS = -nR
ln(Vf/Vi).

Free irreversible expansion:


w=0
If isothermal, ΔU = 0, and therefore q = 0.
Thus, ΔSsur = 0 and ΔStot = ΔS = nR ln (Vf/Vi).

9.22. VARIATION OF ENTROPY WITH TEMPERATURE

The entropy of a system at temperature Tf can be calculated from knowledge of initial temperature
and heat supplied to make ΔT:
𝑓
𝑑𝑞𝑟𝑒𝑣
∆𝑆 = ∫
𝑖 𝑇
When the system is subjected to constant pressure (i.e., the atmosphere) during heating, from the
definition of constant pressure heat capacity, if the system is not doing expansion work (w = 0), then:
dqrev = CpdT
Then, at constant pressure (or constant volume, replace with CV):
𝑓
𝐶𝑝 𝑑𝑇
𝑆(𝑇𝑓 ) = 𝑆(𝑇𝑖 ) + ∫
𝑇
𝑖
If Cp is invariant to temperature change
𝑓
𝐶𝑝 𝑑𝑇 𝑇𝑓
𝑆(𝑇𝑓 ) = 𝑆(𝑇𝑖 ) + ∫ = 𝑆(𝑇𝑖 ) + 𝐶𝑝 𝑙𝑛 ( )
𝑇 𝑇𝑖
𝑖

9.23. CALCULATING THE ENTROPY CHANGE

Calculation of Entropy Changes Accompanying Specific Processes

1. Expansion
The change in entropy of a perfect gas that expands isothermally from V i to Vf is
𝑉𝑓
∆𝑆 = 𝑛𝑅𝑙𝑛
𝑉𝑖
Because S is a state function, the value of ΔS of the system is independent of the path
between the initial and final states, so this expression applies whether the change of state occurs
reversibly or irreversibly.
The logarithmic increase in entropy of a perfect gas as it expands isothermally.
The total change in entropy, however, does not depend on how the expansion takes place.
For any process dqsur = -dq, and for a reversible change we use the expression:
𝑞𝑠𝑢𝑟 𝑞𝑟𝑒𝑣 𝑉𝑓
𝛥𝑆𝑠𝑢𝑟 = = − = −𝑛𝑅𝑙𝑛
𝑇 𝑇 𝑉𝑖
This change is the negative of the change in the system, so we can conclude that ΔS tot = 0,
which is what we should expect for a reversible process. If the isothermal expansion occurs freely (w

RVM Lecture Manual: Physical Chemistry 1 16


= 0) and irreversibly, then q = 0 (because ΔU = 0). Consequently, ΔSsur = 0 and the total entropy
change is given by eq. (4) where in this case, ΔStot > 0 as we expect for an irreversible process.

2. Phase Transition
The degree of dispersal of matter and energy changes when a substance freezes or boils as a
result of changes in the order with which the molecules pack together and the extent to which the
energy is localized or dispersed. Therefore, we should expect the transition to be accompanied by a
change in entropy. For example, when a substance vaporizes, a compact condensed phase changes
into a widely dispersed gas and we can expect the entropy of the substance to increase considerably.
The entropy of a solid also increases when it melts to a liquid and when the liquid turns into gas.
Consider a system and its surroundings at the normal transition temperature, T trs, the
temperature at which two phases are in equilibrium at 1 atm. This temperature is 0°C (273 K) for ice
in equilibrium with liquid water at 1 atm, and 100°C (373 K) for water in equilibrium with its vapour
at 1 atm. At the transition temperature, any transfer of energy as heat between the system and its
surroundings is reversible because the two phases in the system are in equilibrium. Because at
constant pressure, q = ΔtrsH, the change in molar entropy of the system is:
∆𝑡𝑟𝑠 𝐻
∆𝑡𝑟𝑠 𝑆 =
𝑇𝑡𝑟𝑠
If the phase transition is exothermic (ΔtrsH < 0, as in freezing or condensing), then the
entropy change is negative. This decrease in entropy is consistent with localization of matter and
energy that accompanies the formation of a solid from a liquid or a liquid from a gas. If the transition
is endothermic (ΔtrsH > 0, as in melting and vaporization), then the entropy change is positive, which
is consistent with dispersal of energy and matter in the system.

3. Heating
To calculate the entropy of a system at a temperature T f from knowledge of its entropy at a
temperature Ti and the heat supplied to change its temperature from one value to the other:
𝑇𝑓
𝑑𝑞𝑟𝑒𝑣
𝑆(𝑇𝑓 ) = 𝑆(𝑇𝑖 ) + ∫
𝑇𝑖 𝑇
We shall be particularly interested in the entropy change when the system is subjected to
constant pressure (such as from the atmosphere) during the heating. Then, from the definition of
constant-pressure heat capacity, dqrev=CpdT provided the system is doing no non-expansion work.
Consequently, at constant pressure:
𝑇𝑓 𝐶 𝑑𝑇
𝑝
𝑆(𝑇𝑓 ) = 𝑆(𝑇𝑖 ) + ∫
𝑇𝑖 𝑇
The same expression applies at constant volume, but with C p replaced by Cv. When Cp is
independent of temperature in the temperature range of interest, it can be taken outside the integral
and we obtain:
𝑇𝑓
𝑑𝑇 𝑇𝑓
𝑆(𝑇𝑓 ) = 𝑆(𝑇𝑖 ) + ∫ = 𝑆(𝑇𝑖) + 𝐶𝑝𝑙𝑛
𝑇𝑖 𝑇 𝑇𝑖

with a similar expression for heating at constant volume.


The logarithmic increase in entropy of a substance as it is heated at constant volume. Different
curves correspond to different values of the constant-volume heat capacity (which is assumed
constant over the temperature range) expressed as Cv,m/R.

EXAMPLE PROBLEM:

Calculate ΔS when argon at 25oC and 1.00 atm in a container of volume 500 cm3 expands to 1000 cm3 and
is simultaneously heated to 100oC.

RVM Lecture Manual: Physical Chemistry 1 17


SOLUTION:
Methodology: Since S is a state function, we can choose a convenient path from initial to final state: (1)
isothermal expansion to final volume, then (2) reversible heating at constant volume to final temperature.

Amount of Ar present is n = pV/RT = 0.0204 mol


Cp,m(Ar) = 20.786 J K-1 mol-1

(1) Expansion from 500 cm3 to 1000 cm3 at constant T:


ΔS = nR ln 2.00 = +0.118 J K-1

(2) Reversible heating from 25oC to 100oC at constant V:


ΔS = (0.0204 mol) × (12.48 J K-1) × ln (373 K / 298 K)
= +0.057 J K-1
Overall entropy change:
ΔS = +0.118 J K-1 + 0.057 J K-1 = +0.175 J K-1

4. The measurement of entropy


Entropy of a system at temperature T can be related to entropy at T = 0 by measuring heat
capacities at different temperatures, and evaluating
𝑓
𝐶𝑝 𝑑𝑇
𝑆(𝑇𝑓 ) = 𝑆(𝑇𝑖 ) + ∫
𝑇
𝑖
The entropy of transition, ΔtrsS = ΔtrsH/Ttrs, is added for each phase transition between T = 0 and
T (temperature of interest).
For example, if substance melts at Tf and boils at Tb, entropy above Tb is
𝑇𝑓 𝑇𝑏 𝑇
𝐶𝑝 (𝑠)𝑑𝑇 ∆𝑓𝑢𝑠 𝐻 𝐶𝑝 (1)𝑑𝑇 ∆𝑣𝑎𝑝 𝐻 𝐶𝑝 (𝑔)𝑑𝑇
𝑆(𝑇) = 𝑆(𝑂) + ∫ + + ∫ + + ∫
𝑇 𝑇𝑓 𝑇 𝑇𝑏 𝑇
0 𝑇𝑓 𝑇𝑏
All of the quantities can be determined from calorimetry excepting S(0), and integrals can be
evaluated analytically.

Plot (a) shows the variation of Cp/T with sample temperature. The area under the curve of Cp/T as a
function of T is required - since dT/T = d ln T, we can also evaluate area under a plot of Cp vs. ln T.
Plot (b) shows the entropy of the system varying with temperature, which is equal to the area under the
curve up to the corresponding temperature, plus entropy of each phase transition passes.
One problem with measuring ΔS is measuring Cp at low T near T = 0: Debye extrapolation: It has been
shown that at temperatures near T = 0, the heat capacity is approximately equal to T3
(i.e., Cp = aT3 as T 0).

EXAMPLE: CALCULATING ENTROPY

Consider standard molar entropy of N2 (g) at 25oC, calculated from the following data:

RVM Lecture Manual: Physical Chemistry 1 18


Smo (J K-1 mol-1)
Debye exptrapolation 1.92
Integration from 10 - 35.61 K 25.25
Phase transition at 35.61 K 6.43
Integration from 35.61 - 63.14 K 23.38
Fusion at 63.14 K 11.42
Integration from 63.14 - 77.32 K 25.25
Vapourization at 77.32 K 72.13
Integration from 77.32 - 298.15 K 39.20
Correction for Gas Imperfection* 0.92
Total 192.06

Smo (298.15 K) = Smo (0) + 192.1 J K-1 mol-1

CALCULATING ENTROPY AT LOW TEMPERATURE

The molar constant-pressure heat capacity of some solid material at 10 K is 0.43 J K -1 mol-1. What is
the molar entropy at that temperature?
Because temperature is low, we can assume that heat capacity varies with temperature as aT3:
𝑇𝑓 𝑇𝑓
𝑎𝑇 3 𝑑𝑇
𝑆(𝑇) = 𝑆(0) + ∫ = 𝑆(0) + 𝑎 ∫ 𝑇 2 𝑑𝑇
𝑇
0 0
1 1
= 𝑆(0) + 𝑎𝑇 3 = 𝑆(0) + 𝐶𝑝 (𝑇)
3 3
= 𝑆𝑚 (10𝐾) = 𝑆𝑚 (0) + 0.14 𝐽𝐾 −1 𝑚𝑜𝑙 −1
It turns out that the final result can be expressed in terms of heat capacity a constant pressure.

9.24. ENTROPY CHANGES OF MELTING ICE

Consider the entropy changes for putting an ice cube in a glass of warm water and letting it melt
(adiabatic container).

Start at Ti

(1) Calculate ΔS to cool the water to 0oC by reversibly removing heat, q1, from the system.
(2) At Tfus, calculate the amount of heat, q2, to be added to the system to melt the ice cube.
(3) Calculate the difference between the two amounts of heat and add back remaining heat so that the total
heat lost or gained is zero (adiabatic system) - determine the entropy change in this process.
At each step, the infinitesimal entropy change for the system, dS, is just dq divided by the T. For the
cooling and heating of water, integrate over the temperature range, since the temperature is not constant.

Step 1: Entropy decrease for the system as the water is cooled.


𝑇𝑓𝑢𝑠 𝑇𝑓𝑢𝑠
𝑑𝑞 𝑑𝑇 𝑇𝑓𝑢𝑠
∆𝑆1 = ∫ = 𝐶𝑝 ∫ = 𝐶𝑝 𝑙𝑛 ( )
𝑇 𝑇 𝑇𝑖
𝑖 𝑖

RVM Lecture Manual: Physical Chemistry 1 19


Step 2: Melt the ice at the temperature of Tfus = 0oC (qfus = q2).
𝑞𝑓𝑢𝑠 ∆𝐻𝑓𝑢𝑠
∆𝑆2 = =
𝑇 𝑇
Step 3: Balance the heat by (in this example) adding an amount of heat -(q1 + q2) back into the glass (ice
cube melted completely), remove more heat to cool the glass to 0oC than you would have to add in order to
melt the ice -- increase in entropy:
𝑇𝑓
∆𝑆3 = 𝐶𝑝 𝑙𝑛 ( )
𝑇𝑓𝑢𝑠

How do we know Tf?


System is adiabatic so the total heat must be zero, or q3 = -(q1 + q2), giving Tf = Tfus + q3/Cp.

Is the process spontaneous?


(We know intuitively that it is - sticking an ice cube into warm water melts the ice cube!!) How do we
prove this?

Show that the total entropy change (the system plus the surroundings) is positive. The entropy change of the
system we know is:
ΔS1 + ΔS2 + ΔS3

Homework problem 1:
Try the above steps for 5 g of ice at 60oC in 100 mL of water.
With Cp,m of water of 75.6 J K-1 mol-1 and ΔfusH = 6.01 kJ mol-1, Tf should be around 53.4oC with ΔS =
1.39 J K-1 for the system.
Note that since the system is adiabatic, ΔSsur = 0.

Homework problem 2:
Use the same conditions as above, except try the calculation for an isothermal system.
_____________________________________________________________________________________________________________

9.25. USEFUL ENTROPY EQUATIONS

Some changes in state and associated entropy changes for the system, for an infinitessimal change in
entropy, dS = dq/T:
Vary the temperature at constant volume (CV independent of T):
𝑇𝑓 𝑇𝑓
𝑑𝑞 𝐶𝑣 𝑑𝑇 𝑇𝑓
∆𝑆 = ∫ = ∆𝑆 = ∫ = 𝐶𝑣 𝑙𝑛 ( )
𝑇𝑖 𝑇 𝑇𝑖 𝑇 𝑇𝑖
Vary the temperature at constant pressure (Cp independent of T):
𝑇𝑓 𝑇𝑓 𝐶 𝑑𝑇
𝑑𝑞 𝑝 𝑇𝑓
∆𝑆 = ∫ = ∆𝑆 = ∫ = 𝐶𝑝 𝑙𝑛 ( )
𝑇𝑖 𝑇 𝑇𝑖 𝑇 𝑇𝑖
For isothermal expansion of a perfect gas:
𝑉𝑓 𝑃𝑖
∆𝑆 = 𝑛𝑅𝑙𝑛 ( ) = 𝑛𝑅𝑙𝑛 ( )
𝑉𝑖 𝑃𝑓

What if both T and P or T and V change?


Simple, use two steps! First change the T holding either V or P constant, then change V or P at
constant T (depending on what you are given).

_____________________________________________________________________________________________________________

SAMPLE PROBLEM:

RVM Lecture Manual: Physical Chemistry 1 20


1. Calculate the entropy change when argon at 25°C and 1.00 bar in a container of volume 0.500 dm 3 is
allowed to expand to 1.000 dm3 and is simultaneously heated to 100°C.

Solution:
𝑃𝑖 𝑉𝑖
𝑛=
𝑅𝑇𝑖
𝑃𝑖 𝑉𝑖 𝑉𝑓 𝑃𝑖 𝑉𝑖 𝑉𝑓
∆𝑆(𝑆𝑡𝑒𝑝 1) = ( ) 𝑥 𝑅𝑙𝑛 = 𝑙𝑛
𝑅𝑇𝑖 𝑉𝑖 𝑇𝑖 𝑉𝑖
The entropy change in the second step, from 298 K to 373 K at constant volume, is
3
𝑃𝑖 𝑉𝑖 3 𝑇𝑓 𝑃𝑖 𝑉𝑖 𝑇𝑓 2
∆𝑆(𝑆𝑡𝑒𝑝 2) = ( ) 𝑥 𝑅𝑙𝑛 = 𝑙𝑛 ( )
𝑅𝑇𝑖 2 𝑇𝑖 𝑇𝑖 𝑇𝑖
The overall entropy change, the sum of these two changes, is
3 3
𝑃𝑖 𝑉𝑖 𝑉𝑓 𝑃𝑖 𝑉𝑖 𝑇𝑓 2 𝑃𝑖 𝑉𝑖 𝑉𝑓 𝑇𝑓 2
∆𝑆 = 𝑙𝑛 + 𝑙𝑛 ( ) = 𝑙𝑛 { ( ) }
𝑇𝑖 𝑉𝑖 𝑇𝑖 𝑇𝑖 𝑇𝑖 𝑉𝑖 𝑇𝑖

At this point we substitute the data and obtain (by using 1 Pa m 3 = 1J)
3
(1.00 𝑥 105 𝑃𝑎)(0.500 𝑥 10−3 𝑚3 ) 1.000 373 2
∆𝑆 = 𝑙𝑛 { ( ) }
298 𝐾 0.500 298

𝛥𝑆 = 0.173 𝐽𝐾 −1
The laws of thermodynamics we have learned thus far now enable us to begin finding relationships
between properties that may have not been thought to be related with one another - there are many
interesting “hidden” relationships that can be extracted - lets now combine the laws:

The First Law of Thermodynamics:


dU = dq + dw
For reversible change in constant composition system with no non-expansion work
dwrev = -pdV dqrev = TdS

Therefore:
dU = TdS - pdV
Fundamental Equation
dU is an exact differential, independent of path, so the same values for dU are obtained regardless of
change being reversible or irreversible (closed system, no non-expansion work).

9.26. PROPERTIES OF INTERNAL ENERGY

Reversible change: T dS same as dq, and -p dV same as dw


Irreversible change: T dS > dq (Clausius inequality), and -p dV > dw

For a system of constant composition: dw + dq = T dS + -p dV. When S and V are changed, dU α dS


and dU α dV, from the fundamental equation, suggesting that dU should be written as a function of S and V
dU = TdS – pdV
𝔡𝑈 𝔡𝑈
dU = ( 𝔡𝑆 ) dS + (𝔡𝑉 ) dV
𝑉 𝑆
The equation above means that the change in U is proportional to changes in S and V, with the
coefficients being slopes of plots of U against S at constant V, and U against V at constant S. For systems of
constant composition:
𝔡𝑈 𝔡𝑈
Thermodynamic ( ) =T ( ) = -p Thermodynamic
𝔡𝑆 𝑉 𝔡𝑉 𝑆
definition of definition of
temperature temperature

RVM Lecture Manual: Physical Chemistry 1 21


9.27. MAXWELL RELATIONS
The state functions are exact differentials meaning that they must pass the test that indicate their
independence of path taken:
𝔡𝑔 𝔡ℎ
df = g dx + h dy is exact if (𝔡𝑦 ) = (𝔡𝑥 )
𝑥 𝑦
We know that dU = T dS – pdV is exact, then

𝖉𝑻 𝖉𝒑
(𝖉𝑽) = -(𝖉𝑺)
𝑺 𝑽

We have generated a relationship between quantities that would not seem to be related on first sight!
In fact, the four Maxwell relations can be derived from the four state functions U, H, A and G:
𝖉𝑻 𝔡𝑝 𝖉𝑻 𝖉𝑽
From U: ( ) = - ( ) 𝐅𝐫𝐨𝐦 𝐇: ( ) = ( )
𝖉𝑽 𝑺 𝔡𝑆 𝑉 𝖉𝒑 𝑺 𝖉𝑺 𝑷
𝖉𝒑 𝖉𝑺 𝖉𝑽 𝖉𝑺
From A: ( ) = ( ) From G: ( ) = -( )
𝖉𝑻 𝑽 𝖉𝑽 𝑻 𝖉𝑻 𝑷 𝖉𝒑 𝑻

9.28. VARIATION OF INTERNAL ENERGY WITH VOLUME

In Chapter 3, we defined a coefficient called the internal pressure


𝖉𝑼
πT = ( )
𝖉𝑽 𝑻
which can be written as a thermodynamic equation of state, in terms of p and T, and can therefore be
applied to any substance
𝖉𝒑
πT = T ( ) – p
𝖉𝑻 𝑽
This expression for πT is FE obtained by dividing by dV and substituting in:
𝖉𝑼 𝖉𝑼 𝖉𝑼 𝖉𝑼
dU = ( 𝖉𝑺 ) dS + (𝖉𝑽 ) dV ( ) =T
𝖉𝑺 𝑽
( ) = -p
𝖉𝑽 𝑺
𝑽 𝑺
𝖉𝑼 𝖉𝑼 𝖉𝑺 𝖉𝑼
( ) = ( ) ( ) + (𝖉𝑽 )
𝖉𝑽 𝑻 𝖉𝑺 𝑽 𝖉𝑽 𝑻 𝑺
A Maxwell relation
𝖉𝑺 𝖉𝑽 𝖉𝒑
= T (𝖉𝑽) – p (𝖉𝑻) = (𝖉𝑻)
𝑻 𝑷 𝑽

9.29. PROPERTIES OF GIBBS ENERGIES

Same arguments that apply for internal energy, U, are applied to G. When a system has a change of
state resulting in) G, this results from changes in H, S and T. Again, for infinitesimal changes:

dG = dH – TdS – SdT
Since H = U + pV: dH = dU + pdV + Vdp

For a closed system doing no non-expansion work, dU can be replaced by the fundamental equation,

dU = T dS - p dV
dG = (TdS – pdV) + pdV + Vdp – TdS – SdT

thus,
dG = Vdp - SdT

Change in G is proportional to changes in pressure and temperature. Because we control p and T, G is


a very important quantity in chemistry. G carries the combined consequences of the 1st and 2nd laws and we
do not have to worry about ΔSuniv to determine spontaneity!

RVM Lecture Manual: Physical Chemistry 1 22


9.30. GIBBS ENERGY, PRESSURE AND TEMPERATURE
From the relationship below, we can derive two new partial derivatives

dG = Vdp - SdT

𝖉𝑮 𝖉𝑮
( ) = -S ( ) =V
𝖉𝑻 𝑷 𝖉𝒑 𝑻

G decreases when the T is increased at constant p, because S is positive G decreases most sharply
when S is large (G of gaseous phase is more sensitive to) T then liquids or solids)
G increases when p is increased at constant T (since V is positive.

9.31. TEMPERATURE DEPENDENCE OF G

In many chemical and physical transformations, the equilibrium composition of a system depends on
the Gibbs energy - so we must know the response of the Gibbs energy to temperature changes
𝖉𝑮 𝖉𝑮 𝑮−𝑯
Since ( ) = -S we write ( ) =
𝖉𝑻 𝑷 𝖉𝑻 𝑷 𝑻
𝖉𝑮 𝑮 𝑯
And we can rearrange to ( ) - =-
𝖉𝑻 𝑷 𝑻 𝑻
The LHS can be shown to be the derivative
𝔡 𝐺 1 𝖉𝑮 𝒅 𝟏
of G/T wrt T: ( ( )) = ( ) + G
𝔡𝑇 𝑇 𝑇 𝖉𝑻 𝑷 𝒅𝑻 𝑻
𝑃
1 𝖉𝑮 𝑮
(in 59-241, you will see that we want the = ( ) -
𝑇 𝖉𝑻 𝑷 𝑻𝟐
Gibbs-Helmholz equation in this form, since the 1 𝖉𝑮 𝑮
equilibrium constants of reactions are related = [( ) − ]
𝑇 𝖉𝑻 𝑷 𝑻
to G/T rather than
just G) 𝔡 𝐺 𝑯
( ( )) = -
From which follows the Gibbs-Helmholz Equation 𝔡𝑇 𝑇 𝑃 𝑻𝟐
(which relates G and H):
9.32. PRESSURE DEPENDENCE OF G
Calculate the ΔGm of (a) H2O (l) treated as an incompressible fluid, and (b) H2O treated as a perfect
gas, when pressure is increased isothermally from 1.0 bar to 2.0 bar at 298 K. Integrate dG = Vdp - SdT wrt p
at constant T
𝑝
Gm(pf) – Gm(pi) = ∫𝑝 𝑓 𝑉𝑚 𝑑𝑝
𝑖
(a) incompressible fluid, Vm = constant, (b) perfect gas, Vm = RT/p

𝑝
(a) Gm (pf) – Gm (pi) = Vm ∫𝑝 𝑓 𝑑𝑝 = Vm x (pf – pi)
𝑖
= (18.0 × 10-6 m3 mol-1) × (1.0 × 105 Pa) = +1.8 J mol-1

RVM Lecture Manual: Physical Chemistry 1 23


𝑝 𝑅𝑇 𝑝
(b) Gm (pf) – Gm (pi) = ∫𝑝 𝑓 𝑝
𝑑𝑝 = RT ln ( f)
𝑝i
𝑖

= (2.48 kJ mol-1) × ln (2.0) = +1.7 kJ mol-1

ΔG = +ve, though the increase for gas is 1000 times that for liquid

9.33. PRESSURE DEPENDENCE OF G, SOLID & LIQUID

For liquids and solids )V is very small, and in a lot of cases may be neglected (under lab conditions,
Vm)p is small!)

Hence, we can normally assume that Gibbs energies for solids and liquids are independent of
pressure – however, for geophysical problems which involve very high temperatures and pressures, the
volume effect on Gibbs energies cannot be ignored!

Suppose that for a solid phase transition that


ΔtrsV = +1.0 cm3 mol-1, then at a high pressure of 3.0 Mbar (i.e., 3.0 × 10 6 bar), Gibbs energy
of transition from ΔtrsG (1 bar) is calculated as:
ΔtrsG (3 Mbar) = ΔtrsG (1 bar) + ΔVm (pf – pi)
= ΔtrsG (1 bar) + (1.0 x 10-6 m3 mol-1) x (3.0 x 1011 Pa – 1.0 x 105 Pa)
= ΔtrsG (1 bar) + (3.0 x 102 kJ mol-1)

9.34. PRESSURE DEPENDENCE OF G, GASES

RVM Lecture Manual: Physical Chemistry 1 24


Molar volumes of gases are very large, so Gibbs energy may be strongly dependent upon the
pressure.
𝑝 𝑅𝑇 𝑝
Gm (pf) - Gm (pi) = ∫𝑝 f 𝑝
𝑑𝑝 = RT ln ( f)
𝑝i
i

If pressure is increased by 10× at room temperature, then molar Gibbs energy increases by about a
factor of 10. As well, if we set p = po (standard pressure of 1.0 bar), the molar Gibbs energy of perfect gas at
some pressure p is related to its standard value by
𝑝
Gm (p) = 𝐺mo + RT ln (𝑝𝑜)

Chemical Potential

Chemical potential, μ, of a pure substance is defined as


𝔡𝐺
µ = ( )
𝔡𝑛 𝑇,𝑝

Chemical potential is useful for demonstrating how the Gibbs energy of a system changes as a
substance is added to it. For a pure substance, Gibbs energy is G = n × Gm, so
𝔡nGm
µ = ( ) = Gm
𝔡n T,p
Here, the chemical potential is the same as the molar Gibbs energy. For example, chemical potential of
a perfect gas is:
𝑝
µ = µo + RT ln(𝑝𝑜 )

We give it a special name because later we will apply it to mixtures and talk about the chemical
potentials of individual components. However, remember that the chemical potential of some compound is
just its molar Gibbs free energy.
𝑝
µ = µo + RT ln(𝑝𝑜 )

This equation you will use about a hundred times in the next month, so make sure you understand it.
It simply says that we can define a molar free energy for any substance as just the molar free energy
under standard conditions plus some term that depends on the temperature and the natural log of the
relative amount of the substance (expressed as a pressure in the case of a gas or a concentration in the case of
a solute in a liquid).
For an ideal gas, the chemical potential at some pressure is just the chemical potential at the
standard pressure plus the change in entropy associated with changing to the new pressure. In general,
contributions to chemical potential are split into (i) terms that are properties of the molecules in question
and (ii) terms that have to do with changing the total number of states that are available to the molecules.

RVM Lecture Manual: Physical Chemistry 1 25


9.35. REAL GASES & FUGACITY

Consider the pressure dependence of the chemical potential of a


real gas. To adapt to the case of a real gas, we replace the pressure, f,
called the fugacity
𝑓
µ = µo + RT ln ( 𝑜 )
𝑝
Fugacity, from Latin for “fleetness”; refers to the “tendency to
escape”. Fugacity has the same units’ pressure, and is a bit of a “fudge
factor” for treating as real gases. In physical chemistry, since many
properties of materials are derived from chemical potentials, fugacities
are used to describe pressures.
As p → 0, μ coincides with a perfect gas. At intermediate p,
attractive forces dominate (f < p), and at high pressures, repulsion gives
f > p.

9.36. STANDARD STATES OF REAL GASES

A perfect gas is in standard state when pressure is po (1 bar): pressure arises from kinetic energy,
with no interactions taken into account.
The standard state of a real gas is a hypothetical state in which the gas is at po and behaving perfectly.
 We choose a hypothetical standard state in order to standardize the interactions between
gas particles (for different types of gases) by setting all interactions to zero.
 We do not choose the standard state as a gas at pressure approaching zero, since the
chemical potential of a gas will approach negative infinity at this point: (μ → -∞ as p → 0).
 Standard chemical potential of gases, μo, arises solely from internal molecular structure and
properties of gases, and not from the interactions between particles.

9.37. FUGACITY VS. PRESSURE

Fugacity is written as: f = ɸp


where ɸ is the dimensionless fugacity coefficient, which depends on the nature of the gas, the
temperature and the pressure. Thus,
𝑝
µ = µo + RT ln ( 𝑜 ) + RT ln ɸ
𝑝
where μo refers to the standard state gas influenced only by kinetic energy. The second term
describes the perfect gas, and the third term expresses the effect of the molecular forces of a real gas: all gases
become perfect as p approaches 0, so f → p as p → 0, and ɸ → 1 as p → 0.We can prove that the fugacity
coefficient of a gas can be calculated from
𝑃 𝑍−1
ln ɸ = ∫0 ( ) 𝑑𝑝
𝑝
where Z = pV /nRT = pVm/RT is the compression factor of the gas.

9.38. FUGACITY AND COMPRESSION FACTOR, Z


For all gases:
𝑃 𝑓
Gm(p) – Gm(p ʹ) = ∫𝑝ʹ 𝑉m 𝑑𝑝 = µ - µʹ = RT ln (𝑓ʹ)
f is fugacity at pressure p and f ʹ is fugacity at pressure pʹ. Perfect gas:
𝑃 𝑝
∫𝑝ʹ 𝑉perfect,m 𝑑𝑝 = µperfect - µʹperfect = RT ln (𝑝ʹ)
the difference of the equation is
𝑃 𝑓 𝑝
∫𝑝ʹ 𝑉perfect,m 𝑑𝑝 = RT [ln (𝑓ʹ) − ln (𝑝ʹ)]

RVM Lecture Manual: Physical Chemistry 1 26


𝑓 𝑝ʹ 1 𝑃
which is rearranged to: ln ( 𝑥 )= ∫ (𝑉m − 𝑉perfect,m ) 𝑑𝑝
𝑝 𝑓ʹ 𝑅𝑇 0
If p ʹ → 0, the gas behave perfectly, and f ʹ → p ʹ
𝑓 1 𝑃
ln( ) = ln ɸ = ∫ (𝑉m − 𝑉perfect,m ) 𝑑𝑝
𝑝 𝑅𝑇 0
Vperfect, m = RT /p, Vm = RTZ /p, and substituting in gives us proof, q.e.d.

9.39. FUGACITY AND VAN DER WAALS GASES


To evaluate ɸ, experimental data on Z is needed from low pressures up to the pressure of interest -
sometimes, the fugacity can be obtained from the virial coefficients of a gas:
1
ln ɸ = Bʹ p + Cʹ p + …
2

The full VdW equations of state are shown in terms of fugacity coefficients as functions of reduced
pressure (and labelled as curves at reduced temperatures, T/Tc), providing estimates of fugacities for a wide
range of gases.

9.40. FUGACITY SUMMARY

Most gases:
Z < 1 at moderate pressures, attractive forces
Z > 1 at higher pressures, repulsive forces

Z < 1, moderate p Z > 1, high p


𝑃 𝑍−1 𝑃 𝑍−1
ln ɸ = ∫0 ( ) 𝑑𝑝 <0 ln ɸ = ∫0 ( 𝑝
) 𝑑𝑝 > 0
𝑝

f< p f > p
μ < μo μ > μo

particles stick together“escape tendency” lessened particles driven apart “escape tendency” heightened

10.41. FUGACITY CALCULATION


Suppose attractive interactions for ammonia gas can be neglected at 10.00 atm and 298.15 K. Find an
expression for fugacity of a van der Waals gas in terms of the pressure.
We neglect a from the VdW equation (for attractive forces) and write
𝑅𝑇 𝑏𝑝
p=𝑉 Z=1+
m −𝑏 𝑅𝑇

RVM Lecture Manual: Physical Chemistry 1 27


Evaluating the integral for ln ɸ:
𝑃 𝑍−1 𝑃 𝑏 𝑏𝑝
∫0 ( 𝑝
) 𝑑𝑝 = ∫0 ( ) 𝑑𝑝 =
𝑅𝑇 𝑅𝑇
Then solving for f we have
f = pebp/RT
with b = 3.707 × 10-2 L mol-1, then pb/RT = 1.515 × 10-2, giving
f = (10.00 atm) x e0.01515 = 10.2 atm

The repulsive term, b, increases the fugacity above the pressure, so the “escaping tendency” is
greater than if it were perfect.
In Claus, we learned that Vm (s) ≈ Vm (l) is a very good approximation for most cases,
𝔡𝐺
i.e., ( ) = V, dG = VdP @ constant T
𝔡𝑝 𝑇

∫ 𝑑𝐺 = V∫ 𝑑𝑝 = nVm ∫ 𝑑𝑝
ΔG = nVmΔp
However, in some cases, where really enormous pressure changes are made (e.g., geological), the V m
might be pressure dependent.
One such possibility is
V = V(1 atm)(1 – κtp)
Where κT is the isothermal compressibility

RECALL:
−1 𝔡𝑉
κT = ( )
𝑉 𝔡𝑝 𝑇

ΚT is tabulated in the book (tables at the back) so I am not sure why Atkins doesn’t mention this in
the question.
Nonetheless, it is a good exercise which demonstrates a slightly different calculation of ΔG.

EXAMPLE:
Calculate ΔG for 35 g of ethanol (ρ = 0.789 g cm -3) subject to an isothermal pressure increase from 1
atm to 300 atm.
I would include: V varies according to V = V(1 atm)(1 – κTp)
Where κT = 76.8 x 10-6 atm-1

SOLUTION:
dG = Vdp
𝑝
ΔG = ∫𝑝 f 𝑉 (1 − 𝜅 𝑇 𝑝)𝑑𝑝
i
𝑚 𝑝 𝑝
= [∫𝑝 f 𝑑𝑝 − ∫𝑝 f ( 𝜅 𝑇 𝑝)𝑑𝑝]
𝜌 i i
𝑚 𝑝2
=
𝜌
[𝑝 − 𝜅 𝑇
2
]|pipf Note in 6th Ep. Atkins
35 𝑔 𝑥2 integration missed this
=
0.789 𝑥 106
= 4.436 x 10-5 m3 ∫ 𝑥𝑑𝑥 = 2 Atkins missed this
= (4.436 x 10-5 m3) [ 2999 atm – (76.8 x 10-6 atm -1)/2 x {(300 atm)2 – (1 atm)2 }]
= (4.436 x 10-5 m3) [ 2999 atm – 345.6 atm ]
= 0.1177 atm m3 (101325 Pa atm-1)
= 1.193 x 104 J
= +11.9 kJ

RVM Lecture Manual: Physical Chemistry 1 28

You might also like