You are on page 1of 131

An Introduction to Thermodynamics

Classical thermodynamics deals with the flow of energy under conditions of equilibrium or
near-equilibrium and with the associated properties of the equilibrium states of matter. It is a
macroscopic theory, ignoring completely the details of atomic and molecular structure, though not
the existence of atoms and molecules to the extent required for writing chemical reactions. Time
is not recognized as a variable and cannot appear in thermodynamic equations. For students who
have become familiar with atoms and molecules, it may be surprising to find how far one can go
toward treating chemical and physical equilibria without employing any simplified models or
delving into theories of molecular structure.
The detachment of thermodynamics from molecular theory is an important asset. The
fundamental principles of thermodynamics were developed during the 19 th century on the
foundation of two principal axioms, supplemented by a small number of definitions, long before
atomic structure was understood. Because of this lack of dependence of theory on models, even
today we need not worry about our vast ignorance at the molecular level, especially in the areas of
liquids and ionic solutions, in applying thermodynamics to real systems. It has been said, with
some justification, that if you can prove something by thermodynamics you need not do the
experiment. Such a strong statement must be handled with care, but it should become clear in the
following pages that common practice is quite consistent with this assumption.
Two developments associated primarily with the 20 th century introduced substantial new
insights into thermodynamics. Statistical thermodynamics, or statistical physics, originated with
the efforts of Maxwell and of Boltzmann in the late 19 th century and grew with additions by
Gibbs, Planck, Einstein, and many others into a companion science to thermodynamics. Because
statistical thermodynamics relies on specific models of atomic and molecular structure and
interactions, it provides important tests of those models, at the expense of substantially greater
mathematical complexity than classical thermodynamics. More important for present purposes,
statistical mechanics provides much greater insight into the quantities that appear in
thermodynamic equations, and thus a clearer view of why things happen. Thus we will not
hesitate to introduce some basic principles of statistical mechanics (without the extensive
mathematics) when necessary to explain what is going on.
The other new development, largely responsible for the change in physics from what is
generally considered purely Newtonian to relativistic and quantum physics, arose from the
introduction of operational definitions at the end of the 19 th century. This viewpoint requires that
any definition (of energy, position, or time, for example) must include a statement of how we can
measure the quantity. Application of this criterion demands clarification of some quantities that
were introduced casually, without a solid foundation, in the early days of thermodynamics. We
will try to be more careful in explaining what is meant by our symbols, and what can or cannot be
measured, than has been customary in thermodynamic textbooks.
One of the characteristics of thermodynamics is that most of the terms are familiar. Everyone
has heard of energy, of heat, and of work. The difficulty is that we must sharpen our definitions
to distinguish between loosely associated ideas. We will therefore be particularly careful to define
these familiar quantities carefully, often emphasizing what our technical meanings do not include
as much as specifying the intended meanings.

7/10/07 Intro- 1
One of the guidelines of early thermodynamics was that all energy transfers (under
equilibrium conditions) can be classified as either heat transfer of energy because of
temperature differences (thermo) or as work transfer of energy because of forces and
motions (dynamics). It seems appropriate, therefore, to begin with definitions of energy, heat,
and work.

ENERGY. Energy has been a difficult quantity to define because it has so many faces, or forms in
which it may appear. Initially, energy was defined to be the energy of motion, or kinetic energy,
which for most objects under usual conditions is half the mass times the square of the speed,
E = mv2 (1)
The most convenient, and generally reliable, definition of energy is that it is kinetic energy or
any of the other forms of energy which can be changed into kinetic energy or obtained from
kinetic energy. These other forms of energy include rotational energy (a spinning ball or
weathervane), vibrational energy (a mass oscillating up and down on a spring), and potential
energy (a skier at the top of a slope), as well as energy within an object, called internal energy.

HEAT. Most of the internal energy is associated with the nuclei or with the chemical state of the
object. We will generally ignore the nuclear energy. For any given sample of matter, the nuclear
energy typically remains unchanged. Changes of chemical energy will be considered when there
are chemical reactions. For now, we are more concerned with the relatively small portion of the
internal energy that changes when the temperature changes; it is most often called heat, or more
narrowly defined as thermal energy.
The meaning of the thermodynamic term heat can best be explored by consideration of a
few qualitative or semiquantitative experiments. For each of these we will develop a working
hypothesis, select a crucial test, and revise the hypothesis as necessary.
Our understanding of heat is based upon common experiences. When we stand before a fire,
or when we place a pan of water over a gas fire or in contact with an electrically heated coil, our
senses and the change in character of the water tell us that something passes from the fire or hot
coil to nearby objects (specifically to us or to the pan of water). The effect is to heat the
objects, by which we mean that there is a sensation of warmth that can be verified by a
thermometer. The thermometer, in some way, measures this heat. We seek to find the
relationship between temperature, heating, and heat.
Temperature balance? As an initial hypothesis, assume that a thermometer measures the
amount of heat. If so, we should find that a loss of temperature by one body is compensated by a
gain of temperature by another. To test this we put 200 g of hot water, at 90 oC, into each of two
Dewar flasks1 (Figure 1). To the first flask we add 50 g of water initially at 20o C, and stir until

1
These are also called vacuum flasks, because the space between the silvered double walls
has been evacuated, a design developed by Sir James Dewar. Another common name for these
and for containers of different design but for the same purpose is Thermos bottle, which is the
trade name of the American Thermos Products Co.

7/10/07 Intro- 2
the temperature becomes steady. The new temperature is found
to be about 76.0oC. To the second flask we add 25 g of water, at
20oC, and find the final temperature to be about 82.2o C. We must
ask now whether the experiment shows the initial hypothesis to be
fully satisfactory or not.
There has indeed been a loss of temperature by the water in
the flask and a gain in temperature by the water added. But there
is clearly no temperature balance. The water in the flasks
changed temperature only slightly, whereas the water added
increased in temperature several times as much. Also, the water
in the second flask dropped in temperature less than that in the
first flask, but the water added to the second flask increased in
temperature more than that added to the first flask. Examination
of the results (Table 1) shows that the drop in temperature of the
water originally in the flasks is roughly doubled when twice as
much cool water is added.
The experiment just described suggests how the original
hypothesis might be revised. It appears that a larger amount of
water can absorb more heat for a given temperature increase. The
temperature, therefore, is more nearly a concentration of heat.
From this revised hypothesis we predict that the temperature
change times the amount of the substance should be the same for both the added and the original
water (see Table 1).
Table 1 Temperature Measurements*
Sample Tfinal Ts Tw -Tw/Ts -TwMw/TsMs
H2O,50g 76.0 56.0 -14.0 0.25 1.0
H2O,25g 82.2 66.2 -7.8 0.125 1.0
Al, 50g 86.35 66.35 -3.65 0.0050 0.22
Al, 25g 88.12 68.12 -1.88 0.0276 0.22

* Temperatures in oC. Ts is the temperature change of the sample and


Tw is the temperature change of the water originally in the flask.
Initial temperature of the water is 90oC and of the sample, 20o C.
Ms is the mass of the sample added and Mw the mass of water in the flask.

It is necessary to find out whether the same relationship will hold if we exchange heat
between two different substances. To do this we again prepare two flasks, each containing 200 g
of water at 90oC, then add to one a block of aluminum, at 20o C, with a mass of 50 g and to the
second a block of aluminum, at 20oC, with a mass of 25 g (Figure 2). After a few seconds we
may assume that the temperatures of the aluminum blocks are equal to the temperatures of the
surrounding water about 86.35oC for the larger block and 88.12oC for the smaller block.

7/10/07 Intro- 3
Multiplying temperature change by mass and comparing the result for the aluminum block and the
water shows that the ratio is the same for both parts of the experiment with aluminum, but
appreciably different from the results of the earlier experiment. We conclude, therefore, that
aluminum and water have a different heat capacity, so that a given amount of heat added to a
certain mass of one produced a different temperature change than equal heat added to the same
mass of the other.

Volume or mass? We have left unanswered the question whether


the heat capacity depends upon the volume or the mass of the
substance that is absorbing the heat. The choice can be easily made by
means of an experiment employing a substance, such as air, that can
readily change volume without changing mass. We fill one flask with
air, evacuate a second identical flask, and immerse both in water, with
a connection provided between the flasks, as shown in Figure 3. The
temperature of the water is measured; then the stopcock is opened,
allowing the air to expand to twice its initial volume, and the
temperature is remeasured. The temperature is found to be
unchanged. From this we conclude that the temperature of the air did
not change with the change in volume, and therefore that it is better to
define the heat capacity in terms of mass rather than of volume.
(The result is confirmed by more sensitive tests.)
Is heat gained or lost? In each of the measurements described
thus far it has been possible to follow heat as it flows from one body to
another; the amount lost by one substance has been equal to the
amount gained by the other. It is necessary to determine whether this is always true. (If it is, we
would say that heat is conserved, or that the amount of heat is constant.) Taking a hint from
the famous observations of Count Rumford, who noted the great quantities of heat evolved during
the boring of cannons, we design our next experiment to include mechanical motion, in which
energy will be added from motion (i.e., by doing work). Instead of expanding the air from one
flask into an evacuated flask, we can let it expand against a piston, as shown in Figure 4. This
time the temperature of
the gas drops (about
50oC) during the
expansion, even though
we add insulation
around the cylinder to
prevent the flow of
heat outward from the
gas. The change of
temperature cannot be
solely because of the
volume change; the

7/10/07 Intro- 4
previous experiment showed that the change of volume did not cause any change of

A doubling of volume causes a temperature drop


from 25oC to about - 25oC.
temperature. The fact that the gas pushes on the piston, causing it to move, must be the
important difference.
A few additional experiments will provide more information on the relationship between
expansion, with work being done, and temperature effects on gases. (For brevity, only the results
of these experiments will be discussed.) Compression of a gas causes an increase in the
temperature just equal to the decrease of temperature during expansion, if both expansion and
compression processes are slow. It is therefore possible, by repeated expansion and compression,
to cycle the temperature between two values. Any other property of the gas that we might
measure, such as density, volume, or viscosity, will be found to depend only on the temperature as
measured by a thermometer, and not on how that temperature was achieved (for any specified
pressure). In other words, the heating effect of a compression seems to be exactly the same as
the heating effect of a flame or other source of heat. Thus it is possible to compress a gas,
thereby raising its temperature; then extract heat from it by removing the insulation until the gas
has returned to room temperature; expand it into an evacuated space without change of
temperature; compress it to again increase its temperature; extract heat; and so forth, as many
times as we wish.
Clearly, heat is not a quantity that retains its identity after it is absorbed by a substance, for
we can add any amount of heat without changing the properties of a gas in any way (provided
only that the proper amount of work is done by the gas). There is no property that will enable us
to determine the amount of heat added to any substance, or the amount of heat removed. The
description of temperature as the concentration of heat is therefore untenable, and must be
abandoned.
If not heat, then what? Temperature is related to a concentration of something more
fundamental, which can give rise to heat or can cause a gas to do work and which is increased
when the substance absorbs heat or when work is done on the substance. This quantity so
directly related to temperature is called energy.
In classical physics, any measurement of energy is necessarily an energy difference. We often

7/10/07 Intro- 5
write E for energy, where we mean E, the difference between the current energy value and some
implied reference level of energy. There are several meanings for energy. Total energy includes
information on where the sample is located and how it is moving. Measure-ments made on a
sample at rest with respect to our apparatus would measure the internal energy. A small portion
of that internal energy is the thermal energy, which is a collection of several different forms of
energy, or modes of storage of energy, in matter, that can be changed by a change of temperature.
It will be sufficient to let E represent total energy, keeping in mind that the changes we are
primarily concerned with will be changes of internal energy and, most often, changes of thermal
energy.
Transfer of heat. Furthermore, in thermodynamics we are primarily concerned with the
transfer of energy to or from a system, including the transfer of thermal energy. This quantity is
nearly always represented by the symbol Q. Thus Q represents a change of energy of the system; 2

Q = E (2)

To ensure that all the energy transfer occurs as transfer of thermal energy, only, we sharpen
the definition of Q:
Q (thermal energy transfer) is the transfer of energy between two objects in
physical contact as a consequence of a difference in temperature between the
objects.
The importance of this refinement will become apparent later.

WORK. We distinguish work from play; we say that machinery works or doesnt work;
and we read that if a force acts on a body and the body moves through some distance, work is
done on the body. But in thermodynamics, work has a special meaning.
First of all, work is a transfer of energy from one object to another by the action of a force. If
work is done on an object, that increases the energy of the object. 3 Thus, if the only energy
transfer is as work,
W = E (3)

Second, and equally important, all investigations thus far of transfer of energy as work

2
Be particularly careful to distinguish between Q, which is an amount of thermal energy
transferred, and the change in a property, such as thermal energy or E. To write Q displays a
confusion to all who may read your notes.
3
Be aware that older thermodynamic literature quite generally defined work with opposite
sign. Because of an emphasis on steam engines, W was taken to be work done by an object,
rather than work done on an object. Some textbooks have chosen both definitions, changing from
one chapter to another. Again, the amount of energy transferred is W, not W.

7/10/07 Intro- 6
indicate that the amount of work done is equal to a product 4 of the force exerted (on the object)
times the displacement of the point of application of the force.
If more than one force is acting, each force must be multiplied by the distance through which
it acts. Then the work terms may be added.

W = Wi
i =
i
f i dx i (4)

What is the imposter equation? The expression for work done is often abbreviated to work is
equal to force times distance, but this is often interpreted to mean a net force times the distance
traveled by the object. Is this imposter equation ever wrong?


Imposter equation ? Ifnetds =
W=
i ds
f i (5)

Yes, if it is interpreted as giving the work done. We give here just two examples. First, if
you jump, you move upward because the floor exerts a force on you (which is the reaction force
to the force you exert on the floor). You move, and momentum p = (mv) is transferred between
you and the floor, but the force exerted on you times the distance you move (while in contact
with the floor) cannot be equal to work done on you because the floor does not transfer any
energy to you. The point of application of the force is at the floor, and that point doesnt move.
Second, if you drag a box across the floor, it is easy to find the force you exert on the cord
and the distance the cord moves, so work done by you on the cord is known. Similarly, the force
exerted by the cord on the box can be measured, as well as the distance the box moves, so work
done by the cord on the box is known. (Energy transferred, as work, to the cord is equal to
energy transferred, as work, by the cord to the box) But the forces of friction between the box
and the floor are molecular-level forces, not known in detail. And the distances over which these
molecular-level forces act, individually, cannot be known. A simple force times distance
calculation gives an answer, but it is demonstrably wrong! 5
Analogues. Fortunately, we can usually avoid ambiguities in the definition of work in our

4
The product of force and distance is an integral of a scalar product, or dot product.
That will be automatically taken care of in most of our applications. In mechanics, we often
measure work done on a particle (or, better, a physical particle), which can change only its
kinetic energy. For such a particle it is easily shown that work is Ifds. We are concerned here
with more general bodies, subject to rotation and deformation, for which the definition of work
must be more explicit.
5
The imposter equation, (p)2 /2m = Ifnetds, where p = mv, has been called the second-
law equation, SLI. It is a valid statement of Newtons second law, net force is equal to mass
times acceleration, but except for special circumstances it has no connection to W.

7/10/07 Intro- 7
analysis of equilibrium thermodynamics. We must, however, be sensitive to the differences
between what is transferred and what is stored.
The description of energy transfers as heat (thermal energy transfer) and as work may be
compared with deposits and withdrawals from a savings bank. The deposit slip may ask for a
separate listing of bills and coins, and a withdrawal may be in the form of bills or coins. Yet the
account balance itself is neither bills nor coins. In the same manner, energy may be put into a
substance, or withdrawn from it, either as Q, a thermal energy transfer (heat) or as W, work,
but it exists within the substance only as energy not as Q or W. (Remember that money may
also be deposited by check or electronic transfer. We must be alert to the possibility of
transferring energy by means other than Q and W.)
If we are being very casual, it may be sufficient to describe thermal energy and thermal
energy transfers as heat, without additional labels. That, however, is very much like an
accountant choosing a single label (perhaps money) to indicate income and net worth, or trying
to balance your check book without distinguishing between net deposits for the month and
balance at the end of the month. Learning to distinguish between thermal energy and Q, the
transfer of thermal energy, will go a long way toward helping you understand discussions of
thermodynamics.
By going beyond thermodynamics, into statistical mechanics, the internal energy can be
described in terms of thermal energy and other forms of energy (such as chemical), and the
thermal energy may be further broken down into kinetic energy and potential energy of individual
atoms and molecules. But for purposes of thermodynamics we need know nothing more about
energy than that it includes a component related to the temperature. This component, called
thermal energy, can be internally converted to increase or decrease the temperature without
changing the internal energy. (If a mixture of oxygen and hydrogen is ignited, the gases become
substantially hotter, without any transfer of energy to or from the surroundings.) Or the thermal
energy can be transferred to or from the surroundings either as thermal energy transfers or as
work. We can find by experiments, such as the compression-cooling experiment or a variety of
others performed by Joule, how much thermal energy transfer is equivalent to how much work
and, for a given substance, how much energy (by either transfer method) must be put in for a
given temperature rise. The relationships between temperature, energy, thermal energy transfer,
and work will be considered quantitatively in the discussion of the first law of thermodynamics.

7/10/07 Intro- 8
1 The First Law of Thermodynamics
Thermodynamics is based on a small number of postulates, or assumptions. These are called
the laws of thermodynamics because they are suggested by a great amount of accumulated
experimental evidence. In fact it is extremely important to keep in mind that thermodynamics is
important just because there is total agreement between the results of thermodynamics (properly
applied) and all careful experimental results available to us. Because it is not possible to prove
the fundamental assumptions of thermodynamics, both the postulates and the derived results of
thermodynamics have often been challenged. In every showdown thus far, thermodynamics has
been shown to be correct.

Energy
The first step in understanding thermodynamics, and making it serve your purposes, is to
learn how to evaluate changes in energy in any object. As you probably recognize already, that
means understanding how to evaluate Q and W, the transfer of thermal energy to (or from) the
object and the work done on (or by) the object. To do so, we will write what is known as the
first law equation, which directly relates Q and W to the change of energy. But before we can
move ahead, we must look at how we define the object or quantity that is to give up or receive
energy; we must consider what is meant by a property; and we must examine the meaning of a
conservation law.
System and Surroundings. When we consider a ball projected through the air, or a block
sliding down a plane, it is quite satisfactory to talk about the ball, or the block, or simply the
object. If we wish to consider a gas or a solution, object is no longer a very appropriate
description, but we could still refer to the gas or the solution or simply call it it. A better
method is to call whatever is of primary interest (ball, or block, or gas, or solution, or whatever)
the system. One advantage is that we neednt change descriptions when we substitute one gas for
another, or a solution for a gas, and so forth. Another advantage is that it allows for a smooth
transition to problems in which we choose an open system, that is, where we allow material, as
well as energy, to pass back and forth, into or out of the system. Unless specifically indicated
otherwise, however, we will assume our systems to be closed.
Whatever else is around, we simply call the surroundings. Then, ideally, we need consider
only system and surroundings. To avoid difficulties that we might encounter in trying to describe
the far reaches of the universe, it is quite sufficient to limit the term surroundings to include all of
the universe that might be affected by whatever change, or process, we are considering.
It is important that we clearly define what we mean by our system. When we have done so,
the surroundings is usually adequately defined.
There will be times when you will find it more convenient to deal with the interactions of two
bodies, or two systems, with each other, ignoring anything outside those systems. That is
perfectly legal. Terminology such as system and surroundings is meant to be an aid, not a
liability. On the other hand, it has been shown many times that disdaining the definitions of

7/10/07 1- 9
system and surroundings provides sufficient ambiguity to allow erroneous conclusions to be
"proved". Also note that system and surroundings are essentially equivalent and arbitrary labels,
so we are free to interchange them if we wish.
Properties, or State Functions. Examine a block of copper and you will find that it has a
certain shape, a certain volume, a certain temperature, a certain density, and various other
properties that characterize the sample. Each of these may be changed, by changing the
temperature of the copper, or to some extent by changing the pressure exerted on the copper.
Neglecting the shape, which depends on the specific sample but not on the nature of the copper
itself, we call each of these measurable quantities a property of the copper. The state of the
copper, or state of the system, is defined by these properties, so they are often called state
functions.
A gas, such as a gram of oxygen, O2, is adequately described if we know the temperature, the
pressure, and the volume, but two of these are sufficient to determine the third, so only two state
functions are sufficient to define the state of the system. Other substances, such as iron or plastic,
may have additional properties (such as magnetic state or strain) that must be specified.
Conservation Laws. In mechanics we pay special attention to quantities that do not change
during a process we are studying. A free particle (not acted on by any net force) will have a
constant velocity no change in speed or in direction. Such a constant of the motion may be
said to be preserved under the special conditions of the process. In a frictionless system, energy
will be preserved, or unchanged. When a liquid is poured from one container to another, the
volume of the liquid is preserved. Unfortunately, it has become common practice to refer, in such
instances, to the unchanging quality as being conserved, although the term conservation has
quite a different, very important meaning.
Volume is something we can see and often measure with a meter stick, whether the sample is
a copper block or the gram of oxygen gas (in a rigid container). For many processes, the volume
of the system (the liquid) is "preserved", but we also are aware that the volume of such a system
can be changed. With a gas, we need only move a piston, or change the pressure or the
temperature. With the block of copper or a free liquid, a change of temperature will change the
volume.
What we cannot do is change the volume of our selected system without causing some
change elsewhere. If the volume of the system increases, then less volume is available outside the
system, so the volume of the surroundings must decrease, by the same amount. Thus we have the
seemingly trivial, but very important, conclusion that the volume of the system plus the volume of
the surroundings is constant. That, of course, is because of the assumed properties of local space.
We state this result as a conservation law. Volume is a conserved quantity. The volume of a
gas, or liquid, or a solid may be changed, but the volume of the system plus the volume of the
surroundings remains unchanged for all changes we make. Let V = V 2 - V 1 represent the change
in V, the volume. Then the law of conservation of volume would be written

(V )system + surroundings = 0 (1)

The best-known of the three thermodynamic postulates is known as the first law of

7/10/07 1- 10
thermodynamics. It is just the principle of conservation of energy. It may be stated in the form:
First Law The energy of the universe is constant.
A better form for our purposes is to write

First Law (E)system + surroundings = 0 (2)

Remember that (Greek capital delta) indicates a change in the quantity that follows, in this
instance a change in the energy, E. Expressed in symbols,

E / Efinal - Einitial (3)

Because we are primarily interested in changes within that part of the universe that we
choose to call the system, which generally is a small, fixed quantity of substance, a symbol such as
E (unless specifically labeled otherwise) will always refer to the system under discussion. When
the energy of the system increases, E is positive; when the energy of the system decreases, E is
negative. In any given problem, the system to be considered must be carefully defined. The
surroundings are then adequately defined by what is outside the system. It should be clear that
equation 2 is, indeed, a statement of a conservation law, equivalent to the previous statement of
the first law.
It is often convenient to think in terms of a system that is totally isolated from its
surroundings. Unfortunately, such arrangements are not possible. We may prevent energy from
moving to or from a system, or hold volume of the system constant, or keep the pressure on the
system (exerted by the surroundings) constant, or make other special provisions, but there is no
practical way to define a truly isolated system and, if we had one, we could not make
measurements on the system. We must therefore carefully specify in what sense we want our
system to be isolated. It may be isolated in the sense that no energy is transferred to or from the
system, or it may be isolated in a more narrow sense. If the system is insulated, there is no
thermal energy transfer, Q. If the volume of the system is being held constant, there can be no
transfer of energy as work of expansion or compression. To simply call the system isolated
leaves the question indeterminate as to what is being prevented or what is maintained constant.
Heat, Work, and the First-Law Equation. Thus far we have considered energy transfers, to
or from a system, only as Q or as W. These will suffice for the present. Then we can write the
total energy change for the system in the form

E = (E)heat + (E)work (4)

but this is awkward. It is better, and conventional, to introduce new symbols for the terms on
the right-hand side, as we have already indicated. We let Q be the amount of thermal energy
transferred to the system from the surroundings, and let W be the amount of work done on the
system. Then, allowing for both modes of energy transfer, at the same time,

First-Law Equation E = Q + W (4a)

7/10/07 1- 11
Note that this first-law equation, by itself, tells us nothing about whether the energy for the
entire universe changes or not.
Roughly, the division of energy transfers into heat (thermal energy transfer) and work
corresponds to a division into random motion and directed motion (or collective motion). At this
point we have no good criterion for deciding whether light, for example, should be called heat
or work, but for the present it is arbitrary. Later we will find that the second law of
thermodynamics deals with these distinctions.
Insofar as all energy transfers may be measured either as heat (thermal energy transfer) or as
work, or a sum of these, this equation accounts for all possible changes in the energy of the
system. This equation is called the first-law equation. If we combine this equation with the first
law (conservation of energy) we may conclude that Q system = - Q surroundings and that W system = -
Wsurroundings. It is necessary to add additional terms to the first-law equation when there are other
modes of energy transfer (including especially the transfer of matter to or from the system). A
few simple examples of energy changes, in the following sections, will show the meaning of the
terms in equation 4.

TEMPERATURE CHANGES AND HEAT CAPACITY, CV. Consider first the process of heating a gas
confined in a container of constant volume. Because there is no directed motion, no work is
done.

(V) E = Q (5)1

We are often concerned with very small (infinitesimal) changes in energy, which would imply
an infinitesimal value for Q. The temptation is to write dQ, 2 but that would be misunderstood.
Because Q = E, it follows that Q = (E) and dQ = d(E), which is not at all what we
intend. We therefore choose the notation

q = dE (5a)

intended to serve as a reminder that we want an infinitesimal amount of thermal energy transfer.
The amount of thermal energy absorbed by the system, Q, divided by the temperature rise is
the quantity commonly called heat capacity, C. C = Q/T. However, because C may itself vary
with temperature, it is better to take the limiting value of the ratio for small changes of

1
The notation (V) at the left of the equation indicates that the equation is valid when the
volume of the system is held constant. We are also assuming, for the present, that there is no
other form of work, such as electrical work.
2
Other choices are Q and Q, each of which is often misinterpreted, including confusing
the small magnitude with an inexact differential (i.e., change).

7/10/07 1- 12
temperature. This is just a derivative. 3

E Q q
CV = Lim = Lim = (6)
T 0 T T 0 T dT

It is common practice to represent a derivative, such as dE/dT, which is evaluated with some
variable (in this case V) held constant, by the special symbol

E

T V

These are called partial derivatives. Employing this notation we arrive at the usual form of
equation 6.
E
CV = (6a)
T V

Consider now the change in energy for 2 mol of H2 warmed at constant volume from 25 oC to
50 C.4 For a diatomic gas near room temperature, C V is constant and is about 5/2 R = 21
o

J/molK. The gas can do no work at constant volume so the total energy change is equal to the
thermal energy transfer (heat absorbed).

E
(V ) dE =
dE
dT = dT = CV dT (7 )
dT T V
The total energy change is obtained by integration.

(V ) (8)
T2
E = E2 E1 = dE
T1

= CV dT = CV (T2 T1 )
T2

T1

Thermodynamic temperature is expressed on the Kelvin scale, by adding 273.15 to the

3
See Appendix.
4
For convenience, numbers specified for illustrative purposes, such as amounts of
material, temperatures, and pressures, will be treated as integers, known exactly. Also, for
present purposes we approximate CV for hydrogen as 5/2 R (see Table I), although the actual
value is somewhat less and increasing with temperature.

7/10/07 1- 13
temperature on the Celsius5 scale. Temperature differences are the same for the two scales. With
the appropriate numerical substitutions,

E = 2 mol x 21 J/molK x (323 - 298) K


= 1.05 kJ/mol

This is the amount of (thermal) energy that must be added to the gas to raise its temperature
25oC.

IDEAL-GAS EXPANSIONS. Now let 2 mol of hydrogen, at 3 atm pressure, expand at a constant
temperature of 50oC through a pinhole (to maintain a slow expansion) and against a piston.
Assume the external pressure acting on the piston is 1 atm, as in Figure 1. The work done on the
gas is the product of the force exerted on the gas (by the piston) and the distance through which
the piston moves, with a negative sign because the pressure (of the gas) acts opposite to the
pressure acting on the gas. The force, on the gas, times the displacement, is f dx = - (PA) dx = -
P (Adx) = - P dV. Because of the unusual arrangement (effusion controlled), the pressure acting
on the gas at the point of displacement is the external pressure, P ext. Therefore the work done on
the gas is

(Effusion controlled) W = I f dx = - I Pext dV (9)

Because Pext is maintained constant, it may be removed from the integral.

W = - Pext I dV = - Pext V (9a)

5
This scale was formerly called centigrade in English-speaking countries, a name that
can be confused with 1/100 degree in other languages.

7/10/07 1- 14
The initial and final volumes can be calculated from the ideal-gas equation 6, PV = nRT.

2 mol x 8 .314 J/mol K x 323 K


V1 = 5
= 0.0530 m 3
1 . 013 x 10 Pa

2 mol x 8.314 J/mol K x 323 K


V2 = 5
= 0.01767 m 3
3 x 1.013 x 10 Pa

Thus the work is

W = - Pext V = - Pext (Vf - Vi)


= - 1.013 x 105 Pa (0.0530 - 0.01767) m3 = - 3.58 kJ

The necessary constants are given in Table 1.


Table 1 GAS CONSTANT
AND CONVERSION FACTORS

1 joule = 1 newtonmeter = 0.239 cal = 1 pascalmeter 3


1 cal (thermochemical)* = 4.1840 joule = 41.3 mlatm
1 mlatm = 0.1013 joule = 0.0242 cal
1 atm = 1.01325 x 105 N/m2
8.314 J 1.987 cal 82.06 ml atm
R= = =
mol K mol K mol K
* Various experimentally defined values of the calorie appear in
the literature (including the dietician's Calorie = 1 kcal). Only the
thermochemical calorie is defined exactly.

In order that the temperature may remain constant it is necessary to supply thermal energy to
the gas to compensate for the energy expended in doing work. From the first-law equation,

6
The term ideal means simply that the substance obeys a certain equation. An ideal gas
obeys the equation PV = nRT; in later chapters we will encounter the ideal solution, which
obeys an equation known as Raoults law. The ideal-gas equation combines Boyles law and
Charles, or Gay-Lussacs, law into a single, more convenient expression. The temperature must
be on an absolute scale, which we will always take as the Kelvin scale; n is the number of moles of
gas; and R is a universal constant, whose value depends on the units chosen for pressure and
volume. It should be noted that the product of pressure and volume has the dimensions and units
of energy (although PV is not a measure of energy). Such gases as He, H 2, O 2, and N 2 closely
follow the ideal-gas equation at room temperature; such easily condensable gases as CO 2 or H 2O
vapor follow the equation less closely.

7/10/07 1- 15
Q = E - W

For the special case of an ideal gas, the energy depends only on the temperature, not on the
pressure or volume. At constant temperature, therefore, E = 0 and the thermal energy that must
be supplied is

Q = - W = 3.58 kJ

A more important example than the expansion against a constant external pressure is the
reversible and isothermal (constant temperature) expansion. A process is said to be
thermodynamically reversible if it can be reversed at any stage by an infinitesimal increase in the
opposing force or an infinitesimal decrease in the driving force. It should be clear that such a
reversible process is an example of a limiting case, which is only approximately achieved in
practice. Not only does it require a frictionless mechanism, but if the unbalance of forces is
infinitesimal, the rate will be infinitesimal and the time required will be infinite. The question of
reversibility will be considered in more detail later.
From the assumption that the operation is isothermal and the gas is ideal,

E = 0
and
Q = - W = I P dV

Because the expansion is to be reversible, the pressure on the piston differs only infinitesimally
from the equilibrium pressure of the gas, P = nRT/V. The work done on the gas is then

(T, I.G., rev.) W = - I nRT dV/V = - nRT ln(Vf /Vi) (10)

7/10/07 1- 16
For example, if 2 mol of hydrogen at 3 atm and 50oC is expanded isothermally and reversibly
to a final pressure of 1 atm,

Q = - W = 2 mol x 8.314 J/molK x 323 K x ln Vf /Vi

Because the temperature is constant, V f /V i = Pi /Pf = 3.

Q = - W = 5.37 kJ ln 3 = 5.90 kJ

Notice that the amount of thermal energy that must be supplied is greater for the reversible
expansion than in the previous example of expansion against a constant external pressure. A
reversible expansion (or compression) must always give a maximum value for Q, and accordingly
a minimum value for W.7
An integral may be represented by the area under a curve (see Appendix) and the three
examples considered above may be compared graphically. When the gas was warmed, at constant
volume, the pressure increased by the factor T f /T i = 323/297 = 1.09. In a plot of pressure against
volume, this is represented by a vertical line (Figure 2a) and the area under such a line is zero.
No work is performed.
The expansion against a constant opposing force is represented by the plot in Figure 2b. The
gas, in escaping through the pinhole, changes its pressure from 3 atm to 1 atm, but there is no
force exerted on moving surroundings during this escape. When the gas is through the pinhole,
however, it has the pressure of 1 atm and the volume increases, at this pressure, while the gas
pushes against the piston (the surroundings), until all of the gas has reached the same pressure
and occupies the total volume Vf . The work done by the gas, - W, is the area under the vertical
line, which is zero, plus the area under the horizontal line, which is 1 atm times the volume
change.
When the pressure of the gas changes smoothly, as in Figure 2c, the area under the curve (-
W) is a maximum for the given temperature. The pressure follows the curve PV = constant

7
A completely general proof that the reversible work is a minimum for the isothermal
reversible process cannot be given here, but this important result can be illustrated by considering
the expansion or compression of a gas. When the piston moves away from the gas molecules, the
molecules do not strike the piston as hard as when the piston is stationary. The pressure exerted
by the gas on the piston, and hence by the piston on the gas, is therefore less than the equilibrium
pressure of the gas. Hence IPeff dV < I P dV; the work done by the gas is a maximum for a
reversible expansion. Changing the sign, to obtain the work done on the gas, changes the
direction of the inequality; the work done on the gas, W, is a minimum when the effective pressure
is equal to the equilibrium pressure, for expansion or compression.
One may see elsewhere the expression W = - I P ext dV applied when Pext, the external
pressure, differs from the pressure of the system (= the gas). Except in unusual circumstances
(e.g., an effusion-controlled expansion), this is not even a good approximation. See Am. J. Phys.
37, 675-679 (1969).

7/10/07 1- 17
(where the constant is nRT) which is part of a hyperbola.
The condition to be satisfied in order that an expansion or compression of any fluid should be
thermodynamically reversible, and thus that the work done on the fluid should be - IP dV, is
simply that the fluid have a well-defined, uniform pressure throughout. In other words, there
must be no leaks, allowing different pressures at different points, and the motion of the piston
confining the fluid must be slow compared to the relaxation time of the fluid, or the time for
pressures to equilibrate in the fluid. For a gas, the expansion or compression must be very slow
compared to the speed of sound in the gas. Unless there is such a well-defined pressure, or in
certain situations (see Figure 1) two or more well-defined pressures, for the system,
thermodynamics cannot be applied. It is then necessary to apply the more difficult methods of
non-equilibrium, or irreversible, thermodynamics. That is, equilibrium thermodynamics is
sufficient for reversible processes and certain types of irreversible processes, whereas irreversible
thermodynamics deals with the time-dependent equations that become necessary in other
problems involving irreversible phenomena, such as diffusion rates and shock-wave propagation.
The existence throughout the process of a well-defined state (an equilibrium state, for
which pressure is defined) for the system is sufficient to ensure that the expansion or compression
will be reversible, but there may well be other irreversibilities in the total process. Some of the
work done by the gas may be against frictional forces (in the surroundings), or work done on a
massive piston may appear as kinetic energy of the piston, subsequently converted to thermal
energy by collision of the piston with mechanical stops (in the surroundings). Or there may be
conduction of thermal energy to or from the surroundings at a lower or higher temperature. Thus
the total process may consist of a sum of parts, some of which are reversible and some
irreversible. It is particularly important in such circumstances to define carefully the system and
surroundings and the exact process to be considered. Interchanging system and surroundings
labels may be helpful in analyzing the total process.

PHASE CHANGES. Let 5 g of ice melt at 0oC under 1 atm pressure. Experimental measurements
have shown that the amount of thermal energy that must be supplied, called the heat of fusion,
is 334 J/g. The work done is -IP dV = - 1 atm x V. The volumes of ice and water are 1.09
cm3/g and 1.00 cm3/g at 0oC. The energy change for the ice can be calculated from the first-law
equation.
E = Q + W = 5 g x 334 J/g - 1 atm(1.00 - 1.09)ml/g x 5 g
Conversion of the work term from mlatm to J gives

E = 1.67 kJ + 0.046 J = 1.67 kJ

Although the total energy absorbed by the ice, E, is slightly greater than the heat of fusion, Q,
the work done on the ice-water system by the atmosphere (0.05 J) is negligible for nearly all
purposes.
If 5 g of water is vaporized at 100oC and 1 atm pressure, the energy change can be calculated

7/10/07 1- 18
in the same manner. The heat of vaporization is 2.258 kJ/g, 8 and for calculation of the work done
it is sufficient to assume that the water vapor is an ideal gas.

E = Q + W
= 2.258 kJ/g x 5 g - 1 atm[ 5/18 mol x (8.314 J/molK)/1 atm x 373.15 K - 5 x 10 -6 m 3]
= 11.290 kJ - 0.861 kJ = 10.43 kJ

In problems of this type, however, an important shortcut is often satisfactory. The volume of
the liquid is sufficiently small compared to that of the vapor that it may be neglected. Then the
work term becomes
W = - P(Vf - Vi) = - PVf = - PnRT/P =- nRT

The value of R in J/molK may be inserted into this expression to avoid completely the units of
volume and the conversion from atm to Pa. The entire problem can thus be written

E = Q + W = 5 g x 2258 J/g - (5/18) mol x 8.314 J/molK x 373 K


= 11.290 kJ - 0.861 kJ = 10.43 kJ

CHEMICAL REACTIONS. An example of a gas-phase chemical reaction is the combustion of


hydrogen at 100oC and 1 atm pressure (or slightly less) to give water vapor.

2 H2 + O2 &6 2 H2O

The heat of reaction (the amount of thermal energy absorbed by the system) is - 242.5
kJ/mol(H2O). That is, a negative amount of thermal energy is absorbed by the reacting system, or
a positive amount of thermal energy is given up by the reacting system to the surroundings. The
work done on the system is

W = - IP dV = P(Vf - Vi) = - P(nf RT/P - ni RT/P) = - (nf - ni )RT


= - (n)RT (11)
and the energy change of the reacting system is

E = Q + W
= 2 mol(- 242.5 kJ/mol) - (- 1 mol) x 8.314 J/molK x 373 K
= - 481.90 kJ + 3.10 J = - 481.90 kJ/(2 mol H2O)

The same reaction, at 100oC and 1 atm pressure (or slightly more) but producing liquid water
rather than water vapor, will have a heat of reaction that differs from that of the preceding

8
This heat of vaporization is required to separate the molecules from each other that is,
it represents an increase in the potential energy of the molecules. Because the temperature does
not change, the kinetic energy remains constant.

7/10/07 1- 19
problem by the heat of vaporization of water.

Q = - 481.90 kJ - 2 mol x 18 g/mol x 2.258 kJ/g = - 481.90 kJ - 81.2 kJ


= - 563.19 kJ/(2 mol H2O)

The work differs, also, because V is different. Neglecting the volume of the liquid, V =
(n)RT/P = (- 3)RT/P.
W = - P V = 3 RT = 3 mol x 8.314 J/molK x 373K = 9.30 kJ

Then the energy change of the system is

E = Q + W = - 563.19 kJ + 9.30 kJ
= - 553.89 kJ/(2 mol H2O)

Enthalpy
We have covered, in brief outline, the methods of finding the change of energy of any system,
by measuring or calculating the amount of thermal energy transfer, Q, and the amount of work
done, W. But already, we have seen that complications arise, such as having to find work done,
against the atmosphere, when that work term is itself of little interest.
The way to avoid these annoying correction terms, in general, is to define a new quality that
focuses on what we do want to know. Our second step, therefore, in understanding an applying
thermodynamics, is to learn how to define and calculate new variables that will simplify our
measurements and calculations.
Chemical reactions and phase changes are more often carried out at constant pressure than at
constant volume. At constant volume (no work done) the energy change is equal to Q, the
amount of thermal energy transferred to the system, but under constant pressure a correction
must be made for the work performed on the system by the external pressure. Consider a
completely general constant-pressure process in which the only work done is because of the
volume change.9
(P) E = Q + W = Q - P(V2 - V1)

which can be written


E = Q - (P2V2 - P1V1)

where P1 = P2. Set E = E2 - E1 and rearrange.

Q = E2 - E1 + P2V2 - P1V1

9
Later we will consider work done in other ways, especially by electrical fields. We are
explicitly excluding such work terms for the present because they are unnecessary here.

7/10/07 1- 20
= E2 + P2V2 - (E1 + P1V1)
or
(P) Q = (E + PV) (12)

This last equation will be encountered so often that it is a great convenience to introduce a new
symbol for the quantity E + PV.

Definition of Enthalpy H = E + PV (13)

Then, if pressure is constant and the only form of work is W =- IP dV, equation 12 becomes

(P) H = Q (14)

The function H is called the enthalpy.10


Note that the enthalpy is actually defined by equation 13, which contains no restrictions on
pressure, temperature, or volume. Nevertheless, the enthalpy will be found to be most convenient
for problems in which pressure is constant.
A comparison between energy and enthalpy, E and H, can be made by calculating the
enthalpy changes for the same processes for which energy changes were previously found.

TEMPERATURE CHANGES AND HEAT CAPACITY, CP. We found that for 2 mol of H2 warmed at
constant volume from 25oC to 50oC, E = 1.05 kJ. For the same process, the enthalpy change is

H = (E + PV) = E + (PV) = E + V P
= E + V(nRTf /V - nRTi /V) = E + nRT
= E + 2 mol x 8.3144 J/molK x 25 K
= 1.05 kJ + 416 J = 1.47 kJ
The enthalpy change is greater than the energy change because the increase in temperature causes
an increase in P and thus in the product PV.
If the gas is warmed at constant pressure, rather than at constant volume, the enthalpy
change can be calculated in much the same way.

H = E + (PV) = E + P V = E + nR T
= 1.47 kJ

The enthalpy, like the energy, depends only on the temperature, for an ideal gas. It is for this
reason that we find the same H (as well as the same E) when the hydrogen gas is warmed by
25oC whether the process is at constant volume or constant pressure, or under other conditions.

10
As you might guess, H was chosen as a symbol because H = Q, the heat (i.e., the
amount of thermal energy transfer). One must, however, avoid any attempt to associate H with
the (total) thermal energy. Warning: PV has the dimensions, and units, of energy but, as shown
explicitly in equation 13, PV is not energy.

7/10/07 1- 21
It should be observed that in the constant-pressure process the correction term to obtain the
enthalpy change from the energy change is a work term, -W = P V. But in the constant-volume
process the correction term, though of equal magnitude, is not a work term, having instead the
form V P. (You will learn very quickly that I V dP W.)
The constant-pressure warming process can be treated in a somewhat different manner.
Employing equation 14, for constant pressure,

(P) dH = q
Then, dividing by dT,

(P) dH q
=
dT dT

or
H
= CP (15)
T P

For a diatomic gas near room temperature, C P is about 7/2 R = 29 J/molK. Inserting this
(approximate) value for H2,
H
H = dT = CP dT
T P
= 2 mol x 29 J/molK x 25 K
= 1.45 kJ

The heat capacity at constant pressure, C P, is greater than, or occasionally equal to, the heat
capacity at constant volume, CV. The difference is small for solids and liquids (typically about 1.7
J/molK), but for an ideal gas the difference is appreciable and is easily calculated. The constant-
pressure and constant-volume restrictions can be dropped from the derivatives of equations 6a
and 15 for the special case of an ideal gas, because both the energy and enthalpy of an ideal gas
are independent of pressure and volume. Taking 1 mol of gas,

dH dE d (PV ) dE d (RT )
CP = = + = + = CV + R
dT dT dT dT dT

or
(I.G.) CP = CV + R (16)

IDEAL-GAS EXPANSIONS. The work performed on an ideal gas as it expands isothermally was
calculated previously for two important special cases (equations 9 and 10). Because the energy
of the ideal gas is independent of pressure and volume, the energy change must be zero in an
isothermal expansion. Thus we conclude that Q + W = 0; the work done on the gas is equal and

7/10/07 1- 22
opposite to Q, the thermal energy absorbed by the gas in the expansion.
At constant temperature, the product PV is also constant for an ideal gas. Therefore, at
constant temperature, H = (E + PV) = E = 0. Enthalpy change is also zero for any
isothermal expansion (or compression) of an ideal gas.
One often sees the generalization that a gas cools on expansion. That is generally not true
except when the gas does work on its surroundings (thus giving up energy to the surroundings)
with no compensating transfer of thermal energy to the gas (or for certain situations involving
non-ideal gases). Obviously there is no cooling in an isothermal expansion.

PHASE CHANGES. The heat of fusion is the amount of thermal energy absorbed by a solid when it
melts (fuses) under constant pressure. From equation 14 this is identically H for the melting
process. Similarly, the heat of vaporization is H vap. Although Efusion is nearly the same as
Hfusion, because of the small volume changes involved, the difference between Evap and Hvap
can be appreciable, as was shown above. Tabulated values are invariably the enthalpy changes.

CHEMICAL REACTIONS. The heat of reaction, as tabulated, assumes constant pressure and is
therefore identical with the H of reaction. 11 That is, it is assumed that the pressures associated
with each reactant and each product are the same before and after the reaction, so the total
pressure remains constant during the reaction.
The assumption of constant individual pressures is much less important when considering
heats of reaction than in some later applications. Keep in mind, however, in interpreting
thermodynamic processes, that we are not considering a process that starts with reactants, only,
and concludes with products, only. Rather, we assume the only change is the chemical reaction.
Thus we begin with a mixture of products and reactants and allow the reaction to proceed
without change in the physical conditions (temperature, pressure, or concentrations) of the
substances. One way of achieving this experimentally would be to have the total amount of
material large, and allow only a small part of the reactants to combine. (The measured values are
then calculated per mole of a reactant or a product.) We will find, in Chapter 4, an alternative
method is often to measure a reaction in an electrochemical cell, under equilibrium conditions.
We will return to this assumption of fixed conditions when we explicitly consider chemical
equilibria.

State Functions
We have seen that those properties such as energy, enthalpy, temperature, pressure, and
volume that depend on the state of a system are called state functions. If measurements
carried out in Boston, Brooklyn, Birmingham, and Berkeley are to be compared, we must be

11
An exception is that heat of combustion is given as a positive number (although the
amount of thermal energy absorbed by the system is negative) and is therefore - H. If there is
any question concerning the convention followed, look for a reaction such as hydrogen plus
oxygen to give water, where we know energy is given off, so Hreaction is negative.

7/10/07 1- 23
certain that we are dealing with substances in substantially identical states. An important question
is how much information we must have about any system in order to determine its state, and
hence to fix completely its physical and chemical properties.

EXTENSIVE AND INTENSIVE PROPERTIES. If the amount of the system is important, it must be
specified, either by mass or by equivalent measure such as the number of moles. 12 Some
properties, such as energy, heat capacity (C P or C V), or volume, will be proportional to the
amount of material. These are called extensive properties. Other quantities, such as pressure,
temperature, and density, are independent of the amount of material and are called intensive
properties. Properties such as molar volume, density, and molar heat capacity are examples of
intensive properties derived from extensive properties. 13
Apart from the size of the system, the most obvious variable to control is the composition.
For a pure substance, the degree of purity and the nature of the residual impurities should be
known. For a solution, the concentration of each of the components is required (one less
concentration value than the number of components in the solution). Where there are possible
differences of phase (solid, liquid, or vapor, or more than one crystal structure) the phase studied
must be reported. In addition, each worker must be able to reproduce exactly such properties of
the system as temperature, pressure, volume, density, viscosity, and refractive index.
Fortunately, if each system can be assumed to be at equilibrium (in its most stable form, or
state), not all of these variables are required. If the composition and mass are known, it is in
general necessary to know only two additional variables (with occasional others as already
mentioned). For example, if the temperature and pressure are known the volume can be
calculated (assuming the necessary measurements have been previously made) whether the
substance is solid, liquid, or gaseous, and from this the density, viscosity, and refractive index
follow. To define the state of a gas it would be equally satisfactory to specify the pressure and
volume, or the volume and temperature. On the other hand, knowing that 1 g of water has a

12
A mole is a certain number of molecules or other particles (Avogadros number, equal
to 6.02 x 1023). Thus it is necessary to indicate what kind of molecules are being counted (e.g., H
is not the same as H2). In practice, the number of moles is determined from the mass and the
molar mass (or so-called molecular weight) with units of g/mol understood. At one time,
quantities such as gram mole and kilogram mole were defined, but under the modern SI
terminology there is only one mole (symbol mol). Nevertheless, a weight equal to the molar mass
number in pounds may appear under the title of pound mole.
13
Specific quantities are generally defined as dimensionless ratios of the (extensive)
properties of a substance to the (extensive) properties of another substance. Thus specific gravity
is the ratio of the density of a substance to the density of water (or, for gases, the ratio to the
density of air). Specific heat was originally the ratio of heat capacity of a substance to the heat
capacity of water. With Q measured in calories, the heat capacity of water is 1 cal/gK, so the
specific heat was equal to numerical value of the heat capacity in cal/gK. Over time, specific heat
has been taken to be equal to the heat capacity per unit mass (per degree), regardless of the units
of mass or energy. Therefore, units must now be specified.

7/10/07 1- 24
volume of 1.000132 cm3 under 1 atm pressure does not determine whether the temperature is 0o C
or 8.1oC, because the water has a minimum volume, or maximum density, at 4o C. Nor could one
determine the pressure accurately if given the volume and temperature of a liquid or solid. This
demonstrates that some discretion is required when variables other than pressure and temperature
are selected, even if extra variables such as electric or magnetic fields are not relevant.

EXACT DIFFERENTIALS AND LINE INTEGRALS. A property, or state function, can depend only on
the present state of a system. It cannot depend on history that is, on how the system arrived at
that state provided we limit the discussion to equilibrium states, so that there can be no strains
in solids, or other residual effects. It follows, therefore, that the change in such an equilibrium
property, in going from one state to another, can depend only on the initial and final states.
Mathematically, this means that the change, described by an integral, depends only on the limits of
the integral. For example,
E2
E = E 2 E1 = E1
dE

Nothing need be known about the process except the energy of the initial state, E 1, and the
energy of the final state, E2, to find E. Integrands of this type are called exact differentials.
Most integrals we will encounter are well defined in
themselves. For example,

x2 1 2 1 2
x1
xdx =
2
x2
2
x1

(xdy + ydx ) = d ( xy ) = x 2 y 2 x1 y 1
x2 , y2 x2 , y2
x1 , y1
x1 , y1

Some, however, are not defined. For example,

x ,y

2 2
ydx
x1 , y1

cannot be evaluated without additional information. If y is


known as a function of x, then we can make a plot of y vs. x, as
in Figure 3. Then, for example, if y = x2 + 3,

(x )
x2 , y2 x2 1 3 1
ydx = + 3 dx = x 2 + 3 x 2 x 13 + 3 x 1
2
x1 , y1 x1 3 3

7/10/07 1- 25
We can say that Iy dx has been integrated along the line shown in Figure 3. This is called a line
integral.14
Line integrals appear in thermodynamics when it is necessary to evaluate a quantity that
depends on the path. Letting y = P and x = V in the example above gives the familiar integral for
work,

W = -IP dV

It is not sufficient to know initial and final states (the end points of the line); the exact path must
be known. This was demonstrated for the calculation of work in Figure 2. Thermal energy
absorbed by a system also depends on the path, so Q is given by a line integral.
It is worth noting that P dV is a change (in P and V, because P changes with V), so P dV is an
inexact differential. W is an amount, not a change in work or any other property, so it is at best
misleading to describe W as a differential (exact or inexact). Careful terminology can be an aid to
understanding. Careless terminology is a hindrance to understanding.

EQUATION OF STATE. Knowledge of two properties, or state functions, is generally sufficient to


know, in principle, all other state functions (for that equilibrium state, if there are no extraneous
variables such as electric or magnetic fields). For example, if the pressure and temperature of n
moles of an ideal gas are known, the volume can be found from the equation of state, V =
nRT/P, or PV = nRT. Any substance, whether an ideal gas or not, follows an equation of state
relating the pressure, volume, and temperature, but the true equation of state may be a very
complicated function and, indeed, may not be known accurately.
Several functions have been proposed as good approximations to the equations of state for
real gases. The best-known is that suggested by van der Waals:

n2a
P + 2 (V nb ) = nRT
V

In this equation, a and b are constants that depend on the particular gas to be described.
When the volume is large and the temperature is high, the terms involving a and b are negligible
and the gas behaves as an ideal gas; but as the gas is cooled and/or compressed, the behavior
deviates more and more from the ideal-gas law, until eventually the gas condenses to a liquid or
solid.15

14
For a more thorough treatment of exact differentials and line integrals see advanced
calculus books. There is a direct mathematical test to determine whether or not a function of two
variables is an exact differential.
15
The constant b is a measure of the volume physically occupied by the molecules, and
thus not available to the gas. The constant a is best visualized as arising from the curved paths of
the molecules, because of mutual attractions. This decreases the frequency of wall collisions

7/10/07 1- 26
It should also be possible to relate the volume of any liquid or solid to its temperature and
pressure, or to express such other properties as refractive index, heat capacity at constant volume
or pressure, thermal conductivity, heats of vaporization or fusion, or vapor pressures of solids or
liquids, in terms of the temperature and pressure. Some of these equations will be encountered in
later chapters.

Thermochemistry
The application to chemical reactions of the principles developed thus far is called
thermochemistry. In particular, the heats of reaction are measured and tabulated and from these
and from measured heat capacities the enthalpy changes are calculated for other reactions or for
other experimental conditions.

HESSS LAW. The enthalpy change for a chemical reaction, such as the oxidation of sulfur dioxide
to sulfur trioxide

2 SO2 (g) + O2 &6 2 SO3 (liq)

can be expressed as the difference between the enthalpies of the initial and final states.

Hreaction = Hfinal - Hinitial


= H(2 SO3) - H(2 SO2) - H(O2)

There is no way within thermodynamics of measuring an absolute energy, 16 or an absolute


enthalpy. Only energy, and enthalpy, changes can be determined. However, knowing that these
energy and enthalpy changes depend only on the initial and final states, it is possible to add and
subtract chemical reactions and add and subtract the corresponding enthalpy changes. That is, we
may quite arbitrarily select a reference energy and/or enthalpy level and measure all values from
that arbitrary level. In particular, it is possible to tabulate heats of formation, the enthalpy
changes in the reaction of the elements to form each compound, and from these to calculate
enthalpies of other reactions. This principle is known as Hesss law.
The reactions for the formation of the gases SO 2 and SO3 from the elements are

S + O2 ----- SO2
S + 3/2 O2 ----- SO3 (liq)

without changing their effect, and therefore decreases the pressure on the walls, requiring the
supplement to P to fit the ideal gas equation form; see Phys. Teach. 34, 248-249 (April, 1996).
16
The reader with some knowledge of special relativity may recognize that the total
energy of any system is measured by its mass, multiplied by the square of the speed of light.
However, it would be necessary to measure masses about a million times more accurately than is
now possible to be able to determine energies to the accuracy required in thermochemistry.

7/10/07 1- 27
The measured enthalpy changes for these reactions at 25 oC and 1 atm pressure are -296.90 kJ/mol
and -437.94 kJ/mol. Subtraction of the first reaction from the second gives
SO2 + O2 &6 SO3 (liq)

and subtraction of the enthalpy changes gives -141.04 kJ/mol, which is the heat of reaction for the
oxidation of SO2 to SO3 (liq).
Exactly the same elements, in the same quantities, always appear on both sides of a chemical
equation (which is why reactions as written are called equations). Subtraction of the elements
from both sides of an equation will yield, on each side, product minus reactants for the reactions
of formation of each of the substances appearing in the original equation. In the example above,
the original equation was SO2 + O2 &6 SO3. Subtract 1 mol of S and 3/2 mol of O2 from each
side. The equation can then be written as the formation of each compound (i.e., of SO 2, O 2, and
SO3) from the elements.
(SO2 - S - O2) + ( O2 - O2) &6 (SO3 - S - 3/2 O2 )
and therefore
Hreaction = Hform(SO3) - Hform(SO2) - Hform( O2)
= -437.94 kJ/mol -296.90 kJ/mol - 0
= -141.04 kJ/mol(SO3 liq)

(Notice that the heat of formation of any element, in its standard state, is necessarily zero.)
An entirely equivalent way of obtaining the same numbers is to consider the enthalpy of each
compound on a scale taken with reference to the elements. Such enthalpy values are called
standard enthalpies of the compounds; they are identical with the standard enthalpies of
formation.
Hesss law can often be applied to find heats of reaction that could not be directly measured
experimentally. For example, the reaction of two molecules of ethylene, C 2H 4, to form
cyclobutane, C4H8, would not readily occur quantitatively under conditions conducive to
measurement of the heat of reaction. But both ethylene and cyclobutane can be burned in
oxygen, and subtraction of these reactions gives the reaction equation desired.

2 C2H4 + 8 O2 &6 4 CO2 + 4 H2O


C4H8 + 8 O2 &6 4 CO2 + 4 H2O

Subtraction of the second from the first gives


2 C2H4 &6 C4H8

and, therefore, subtraction of the H for the second combustion from the H for the first
combustion gives H for the condensation reaction. Heats of combustion (equal to - H reaction )
are comparatively easy to measure and are often tabulated.

KIRCHHOFFS LAW. The heat of reaction at a temperature other than that given in a table can be

7/10/07 1- 28
found by calculating enthalpy changes along an arbitrary path. The total enthalpy change is
independent of this choice of path. The method is known as Kirchhoffs law.
Assume that H is known for a reaction at a temperature T 1 and the H at another temperature,
T2, is to be found. Starting with the hot reactants at T2 (Figure 4), the reaction could be carried
out isothermally to obtain products at the same temperature. An alternative path would be to
cool the reactants to the temperature T 1, carry out the reaction isothermally at T1 , and warm the
products to T2. The heat of reaction at T1 is already known and if the heat capacities at constant
pressure are known, the enthalpy changes can be calculated for the processes of cooling reactants
and warming products. This path must give the same H as the isothermal reaction at T 2.
T1 T2
H 2 = T2
C P ( reactants ) dT + H 1 + T1
C P ( products ) dT (17)
or,
because interchanging limits of an integral will change the sign,

[C ] dT
T2
H 2 = H 1 + P ( products ) C P ( reactants )
T1

If the difference in heat capacities is independent of temperature, this may be rewritten in the form
H2 = H1 + [Cp (products) - CP (reactants)](T2 - T1 ) (18)

For example, given that the heat of reaction for rhombic sulfur burning in oxygen to yield
sulfur dioxide gas is - 296.9 kJ/mol at 25 oC (298 K), find H at 95oC (368 K). The heat
capacities are given in Table 2. Insertion of the numerical values into equation 18 gives

Table 2 HEAT CAPACITIES '


Average values (in J/mol-K) for temperature ranges indicated
Compound CP Temperature, oC
He 20.8 -200 up
H2 28.8 25 to 200
O2 29.4 25 to 200
H2O(g)* 36.4 25 to 200
SO2(g) 41.9 25 to 200
S (r) 23.7 25 to 200
S (m) 25.9 95 to 120
*For rough calculations it is sufficient to set
CP(steam) = CP(ice) = CP(liq H2O).

7/10/07 1- 29
H368 = - 296,900 J/mol + (41.9- 29.4 - 23.7) x 70 J/mol
= - 296.1 J/mol

Sometimes there will be a phase transition during the warming or cooling process. Sulfur has
a phase change at 95oC, at which point rhombic sulfur goes to monoclinic sulfur; the monoclinic
sulfur melts at 119oC. The enthalpy changes are 11.78 and 39.24 kJ/mol. The heat of reaction
for liquid sulfur burning in oxygen to form SO2 at 119oC (392 K) can be calculated as follows (see
Figure 5).
H392 = - Hfusion - CP (m) (119 - 95) - Htr - CP (r) (95 - 25) - CP (O2 ) (119 - 25)
+ H298 + CP (SO2) (119 - 25)
H392 = - 39,240 - 25.9 x 24 - 11,780 - 23.7 x 70 - 29.4 x 94 - 296,900 + 41.9 x 94 J/mol
= - 349.0 kJ/mol
Note that temperature differences can be found without conversion to the Kelvin scale.
Both Hesss law and Kirchhoffs law are simply applications of the principle that changes in

a state function, such as the enthalpy, are completely determined by the initial and final states. 17
This principle is combined with the equation arising from the first law that shows that i f the
pressure is constant, the enthalpy change will be equal to the heat absorbed by the system. Thus
the heat of reaction, by which we mean Hreaction (at a particular temperature, pressure, and
concentrations of reactants and products), is only equal to the heat absorbed i f the reaction
proceeds at constant pressure (and at the specified temperature and concentrations). It is
sometimes more convenient to carry out a reaction at constant volume. Then the heat absorbed is
not equal to the heat of reaction (that is, to H reaction), but it is still determinate because heat
absorbed equals the change in energy when the system follows a constant-volume path and
because Ereaction is fixed by the initial and final states.

17
This property associated with state functions has sometimes been confused with a
conservation principle. Enthalpy is not conserved.

7/10/07 1- 30
The experimental determination of a heat of reaction is called calorimetry. A typical
calorimeter (Figure 6) consists of a reaction chamber, surrounded by a layer of water, enclosed
by sufficient insulation to prevent heat loss to the surrounding laboratory. The reactants, at room
temperature, are placed in the reaction chamber, the calorimeter is closed, and the reaction is
initiated by an electrically heated wire or other controlled energy source. The reaction will
normally be exothermic and the reaction chamber will therefore become quite hot, but the heat is
conducted into the surrounding water layer so that the products, and the water, reach a final
temperature only slightly above the initial temperature. The E reaction is the same as if the entire
process had occurred at the initial temperature even though the materials may have become quite
hot during the course of the reaction. The heat given off by the reaction is calculated by
observing the temperature rise of the water, using the condition that all heat given off by the
reaction must have been absorbed by the water. Small corrections are required for the change of
temperature, from the initial room temperature, of the products, and for the small amount of
energy added by the hot wire or other initiation method. Some systems may also require a
correction for changes of concentrations during the reaction.

7/10/07 1- 31
Problems
1. At room temperature the heat capacity (C V) of most solid elements (except the very light ones)
is about 3R. Specific heat is the heat capacity per unit mass, or the value in J/gK. (For solids the
difference between constant volume and constant pressure conditions is small and may be
neglected here.)
a. Find the specific heat of Pb (molar mass = 207.19).
b. Find the specific heat of Cu (molar mass = 63.546).
c. Find E for 25 g of Cu when it is warmed from 20oC to 35oC.
2. Estimate the final temperature when 10 cm 3 of iron (density 7.86 g/cm3) at 80oC is added to 30
ml of water at 20oC.
3. Find the work done when 3 mol of N2 gas at 4 atm pressure expands slowly at 25 oC against a
constant pressure of 1 atm.
4. Calculate the work done when 0.25 mol of SO2 at 27oC and 1 atm expands reversibly and
isothermally to a final pressure of 0.20 atm. Find Q for the gas during this process.
5. Calculate the final temperature if an ice cube (25 g) at - 5 oC is added to a cup of coffee (150
ml, or 2/3 cup) at 90oC, neglecting heat loss to the surroundings.
6. The heat of vaporization of benzene, C6H6, is 30.72 kJ/mol and the normal boiling point is
80.1oC.
a. What is Evap if the vapor is an ideal gas?
b. Part of the thermal energy absorbed by a liquid as it vaporizes to a gas is required to
perform work on the atmosphere as the substance expands to a vapor. What fraction of the total
Q goes into work against the atmosphere when benzene (density about 0.88 g/cm 3) vaporizes at
its normal boiling point?
7. Find E at 25oC for the oxidation of S (rhombic) to give SO2 (gas) and for the oxidation of S
to give SO3 (liq). The heats of reaction are - 296.81 and - 441.0 kJ/mol. Assume gases are ideal.
8. Calculate H and E for the reaction, at 25oC,

SO2 (g) + O2 ----6 SO3 (g)

The heat of vaporization of SO3 at 25oC is 43.14 kJ/mol.


9. Manganese can be prepared by a thermite process,

3 Mn3O4 + 8 Al -----6 9 Mn + 4 Al2 O3

The standard enthalpy of formation of Mn3O4 is - 1387.8 kJ/mol and for Al2O3, - 1657.7 kJ/mol.
Find the amount of thermal energy given off by the reaction as written above, starting with the
reactants at room temperature and ending with the products at room temperature.
10. The interconversion of graphite and diamond does not occur at room temperature. Explain
how you could determine H for this transition by measurements in the laboratory.
11. Calculate the heat of vaporization of water at 50 oC.
12. Find H for the hydrogenation of ethylene to produce ethane,

C2H4 + H2 -----6 C2H6

7/10/07 1- 32
at 150oC. The standard heats of formation, at 25o C, of ethylene and ethane are 52.4 and - 84.0
kJ/mol, respectively. Heat capacities (C P) of ethylene, hydrogen, and ethane within this
temperature interval are approximately 42.9, 28.8, and 52.5 J/molK. All three compounds are
gases at room temperature and above.
13. Benzene, C6H6, is reduced commercially with hydrogen to give cyclohexane, C6 H12 .

C6H6 + 3 H2 -----6 C6H12

At room temperature (25oC) the benzene and cyclohexane are liquids and the heat of reaction is
- 205.5 kJ/mol. Benzene boils at 80.1oC with a heat of vaporization of 47.8 kJ/mol; cyclohexane
boils at 80.7oC with a heat of vaporization of 33.0 kJ/mol. The average heat capacity of liquid
benzene is 136.0 J/molK, of liquid cyclohexane 154.9 J/molK, of benzene vapor 82.4 J/molK,
of cyclohexane vapor 150 J/molK, and of hydrogen 28.8 J/molK. Find H for the reduction of
benzene with hydrogen at 150oC.
[0.44 cal/gK; 0.41; 26.7 35.8 cal/molK; 6.94 cal/molK]Check numbers
14. Air is approximately 20% O2 and 80% N2, by volume. Find the effective molar mass of air.
Calculate the density of air at 25oC and 1 atm in kg/m3, assuming it acts as an ideal gas.
15. A gaseous compound, containing only sulfur and fluorine, is 62.7% (by weight) sulfur. At
27oC and 750 torr18 the density of the gas is 4.09 g/L. What is the molecular formula of the
compound? Assume the gas is ideal.
16. Find the pressure exerted by 32 g of methane in a 2 L steel bomb at 127 oC, assuming the gas
is ideal.
17. Find the pressure exerted by 32 g of methane in a 2 L steel bomb at 127 oC assuming the gas
obeys van der Waals equation. The constants may be taken as a = 2.25 L 2atm/mol 2 and b =
0.0428 L/mol.
18. Find the volume occupied by 32 g of methane at a pressure of 5 atm at 27 oC if the gas is
ideal. Calculate, with this value, the n2a/V2 correction term in the van der Waals equation and
find, from this, an approximate value for the volume occupied by the methane if it obeys van der
Waals equation. Recalculate the correction term, with this better volume, and recalculate the
volume if necessary.
19. A good vacuum obtained with a mechanical pump and a mercury diffusion pump will
measure 10-5 torr, or better, of noncondensable gases (such as air), but there is 10-3 torr of
mercury vapor present unless this has been trapped out. Very good vacuum systems can give 10 -
9
torr.
a. How many molecules/cm3 are there at a pressure of 10-3 torr?
b. How many molecules/cm3 are there at a pressure of 10-9 torr?
c. What pressure would be required to achieve 1 molecule/cm 3?

18
The torr is a unit of pressure equal to 1/760 atm. It is numerically equivalent to the unit
mm of Hg but is less awkward, especially when mercury is one of several vapors present. The
torr is named after Evangelista Torricelli, who invented the barometer in the 17 th century.

7/10/07 2- 33
2 The Second Law of Thermodynamics
During the 19th century several men, of quite different backgrounds and interests, struggled
with the basic problems of thermodynamics. Brilliant flashes of understanding were followed by
years of doubting, testing, and interpreting. Among the fundamental difficulties was a confusion
between two concepts. One of these was the concept of energy. The other, which is related to
the general concept of equilibrium and the direction of changes with respect to equilibrium, was
stated first, but the language was such that it was long misunderstood and therefore rejected. It is
now accepted as the second law of thermodynamics.
Very few people today who have any acquaintance with modern science would doubt the
following generalizations:
1. Perpetual-motion machines dont work.
2. Bodies in equilibrium have the same temperature. When two bodies in contact have
different temperatures, energy flows, as heat (Q), from the warmer to the cooler body.
3. Bodies in equilibrium have the same pressure. When two bodies in contact, at the same
temperature, have different pressures, the body at the higher pressure tends to expand and
compress the body at the lower pressure.
None of these statements is required by the first law, 1 which requires energy balance in any
process but does not say whether a process will actually occur. For example, both exothermic
and endothermic reactions are known that do proceed without external forcing. There is, despite
the variety in these generalizations, a similarity of pattern and intent saying, for example,
whether a specific process will or will not occur that suggests a common basis. The basis is
found in the postulate known as the second law of thermodynamics.

The Spread Function


We have seen that thermodynamics largely ignores the atomic structure of matter. It is quite
sufficient, for almost all purposes, to measure the macroscopic (large scale) variables such as
temperature, pressure, volume, and energy. The difficulty with this approach is that it effectively
hides from us one key property that may be called the spread function, S.
Consider the thought experiment of adding 10 grains of yellow rice to a large bag of white
rice, which will subsequently be mixed and poured and just thoroughly scrambled. Initially the
yellow grains were carefully arranged, on top in one corner. Where will they be a little later after
mixing?
There is no energy advantage for the yellow grains to move to the bottom, or back to the
top, or to any other specific location. We know that, once mixed, there is close to zero

1
The first law prohibits certain perpetual-motion machines, called perpetual-motion
machines of the first kind. The more interesting attempts, called perpetual-motion machines of
the second kind, do satisfy the first law.

7/10/07 2- 34
probability that they will spontaneously reassemble. Even if the bag of rice is only 100 grains
deep, 100 grains wide, and 100 grains in the third direction, each yellow grain has 100 x 100 x
100 = 106 possible locations, which means that two grains have 106 x 106 = 1012 possible locations
and the 10 grains have (106)10 = 1060 possible locations. The probability they will all collect in
one location is not much more than 1 in 1060, which is awfully close to zero for most purposes.2
Now imagine a very small number of molecules (say 10 6) to which we add 10 molecules that
are excited perhaps they are each vibrating, or rotating, or otherwise have some extra energy
that can be transferred to a neighboring molecule. After a very short time, the extra energy will
be spread throughout the 106 molecules, so there is no significant chance the energy will ever
again (in the lifetime of the universe) be collected in the original 10 molecules. With a more
typical number of molecules (such as 6 x 10 23) and many ways for each to store energy, it is not at
all unreasonable to say the probability of unspreading the energy goes quickly to zero.
If we have a small steel ball (which will typically have 10 23 or more atoms), one way of
storing energy in the ball is to start it rolling. Another very large group of ways of storing energy
in the ball is to let some of the atoms begin to vibrate. The mechanical energy of rolling may be
converted by rolling friction to thermal energy of internal motions. There are so many possible
internal motion states that we dont expect the energy to ever spontaneously reappear as
mechanical energy of rolling.

Entropy
Fortunately, there is an easy way to keep track of S, the spread function, without even having
to count molecules. Any energy passing into an object by thermal energy transfer, Q, is gone in
terms of having been spread around. It shows up afterward only as an increase (perhaps slight) in
the temperature of the system. So the first step is to keep track of Q, and in particular, we want
Q for a reversible path, or one that is as close as possible to equilibrium.
Then we want to know whether this amount of energy is important to the system, or whether
there is already so much energy spread around that this addition is relatively unimportant. That
average energy we will find is nicely measured by the (absolute) temperature of the system. So
we measure the increase in average energy as Q rev and divide by the temperature, T.
Q rev
Definition of Entropy Change S = (1)
T

2
If the ten yellow grains were regarded as indistinguishable, we should remove 10! = 3.6 x
6
10 possibilities that only interchange yellow grains among themselves. That would still leave
more than 1053 possibilities, scarcely different qualitatively from 1060 . The problem of overlapping
positions (two yellow grains being assigned to the same location) may safely be ignored at this
level of probability.

7/10/07 2- 35
The official name of the spread function, S, is entropy 3. By the methods of statistical
mechanics we can show that S is a measure of the actual number of individual states over which
the energy of the system is spread. The greater the number of states (hence S), the greater the
probability associated with the overall state.
It is, of course, possible to transfer energy to our system without increasing the spread of
energy. For example, we can transfer just enough energy, in the proper form, to make the steel
ball start rolling. In a reversible process, this energy transfer is measured as work. As we saw for
ideal gas expansions and compressions, the work done on the system is a minimum (and Q is a
maximum) when the process is reversible. When the process is purely one of transferring energy
to a mechanical mode, such as rolling or moving in a straight line, the amount of thermal energy
transfer, Q = Qmax, is equal to zero. There is no increase in entropy.
An alternative way of approaching the second law is to recognize that E is independent of
path, so we can write, for any process at constant temperature, 4 between fixed end points

(T) E = Q + W = Qrev + Wrev = Qmax + Wmin (2)

There can be only one value of Qmax and one value of Wmin, so we can represent these, also, as
values independent of path, which for the moment we will label as A (= W rev) and as B (=
Qrev) and regard A and B as state functions.
(T) E = A + B
or
E=A+B (3)

In any isothermal process in which the energy of the system changes by E, some part of this
energy change, equal to A, must appear as work done on the system. The second energy term,
called B for the moment, cannot be lost, because of the first law (conservation of energy);
however, it may be regarded as spilled energy, which is lost for useful purposes during the
transfer operation because it has become spread through the system.
From equation 3, two important functions are obtained. One, temporarily given the symbol
A, is called the Helmholtz free energy, or sometimes the work function, and is characterized, as
above, by the constant temperature equation

3
The name was chosen by Clausius from the Greek verb entrepo (), meaning to
turn, implying change, and because it somewhat resembles the word energy, one of the
important quantities involved in finding the number of states Entropy, S, is proportional to (or for
suitable units, equal to) the logarithm of the relative probability, i.e., to the number of states
accessible to the system. Unfortunately, it is also vaguely suggestive of enthalpy, H, which has
different dimensions, units, and meaning.
4
We also assume here that the process satisfies the first-law equation, E = Q + W.
However, equations 1 and 5, and equation 6 (obtained below), are not subject to this restriction
against other forms of work.

7/10/07 2- 36
(T) A = Wrev (4)

This function (which we will subsequently designate by the more current notation, F), and an
even more useful quantity derived from it, will be investigated later.
The spread function, S, is obtained from the second function, B, when we divide by the
temperature, T. That is, S = B/T or B = TS, where S is the spread function, entropy, previously
introduced. Then a small change in B breaks into two parts.

d(TS) = T dS + S dT
and at constant temperature

(T) d(TS) = T dS = qrev

or, quite generally,


q rev
dS = (5)
T

Statistical mechanics tells us something about the value of S, representing the number of
molecular states, but thermodynamics (which never quite admits the importance of molecules)
tells us only how to find changes in entropy, S.

ENTROPY, EQUILIBRIUM, AND SPONTANEOUS CHANGE. Consider the process of melting 1 g of


ice, at 0oC and 1 atm pressure, by bringing it into contact with a thermostat, or heat reservoir.
The thermostat might be a very large block of metal, well insulated from its surroundings and
large enough so that the reasonable amount of thermal energy added or withdrawn will not
change its temperature. For the ice,
Q rev 333 .6 J/g
S = = = 1 .22 J/g K
T 273 .15 K

If the thermostat is at exactly the same temperature,


333.6 J/g (ice)
Sthermostat = = 1.22 J/g K
273.15 K

Taking the ice as the system and the thermostat as the surroundings, (S) system + surroundings = 0. Thus
far it might appear as if entropy also is conserved (compare equation 1, Chapter 1). But we know
that unless there is at least a small temperature differential between the thermostat and the ice, the
ice will not melt. We cannot melt an ice cube by bringing it into contact with a large block of ice;
it must be touched to something warmer. If the temperature of the thermostat is higher than
273.15 K, Sthermostat will be smaller in magnitude than Sice , and S for the ice plus the
thermostat, or system plus surroundings, will be positive. This is the essence of the second law of
thermodynamics, which states

7/10/07 2- 37
(S)system + surroundings > 0 (6)

for any process whatever that actually occurs. The similarity, and contrast, to the statement of
the first law (equation 2, Chapter 1) should be noted. In the limit of a process that is truly
reversible (for example, thermostat and ice at the same temperature, so that the ice can melt or
refreeze by exchange of thermal energy with the thermostat), the total entropy change will
approach the value zero. Recognizing the importance of this limiting condition we may write

Second Law (S)system + surroundings $ 0 (6a)

with the understanding that the equality can be approached as closely as we wish, if we have
sufficient patience.
From equation 6 or 6a, taken as the second basic postulate, it is possible to derive results that
are well established from our experience. There are, in general, many different paths that a
system may follow between any two specified states, of which at least one (actually, an infinite
number) will be thermodynamically reversible. Let q i and w i be the amount of thermal energy
transferred and the amount of work done along the i th path, or more generally, q 1, q 2, q rev are
the various amounts of thermal energy transferred and w 1, w 2, w rev the various amounts of
work done along these paths. Then, in general, q1 q2 qrev and w1 w2 wrev. But dE
is the same for each of these paths. That is, regardless of which path the system actually follows,
dE = q1 + w1 = q2 + w2 = = qrev + wrev

and, because qrev = T dS and wrev = - P dV (if the only work is work of expansion or compression),
it follows that, for any path,

dE = T dS - P dV (7)

This equation is valid for any process occurring in a system of constant composition. (If there is a
change of chemical composition, as would be required if electrical work were produced, or a
change of concentrations, additional terms should be added to the first-law equation and to
equation 7.) At constant volume the second term of equation 7 becomes zero and we obtain

E
=T (8)
S V
or
S 1
= (8a)
E V T

7/10/07 2- 38
Take an arbitrary system, at constant volume, that
is thermally isolated from its surroundings (so that Q =
0 and E = 0). The second law then requires that dS $
0 for any changes within this system. Divide the system
into two parts in any fashion (Figure 1). For example,
the total system might be a cup of water, which is
considered to be divided into two equal parts, or it
might be a bar of lead, with one cm3 in the center
considered as one part and the remainder as the second
part. Let E1 be the energy of the first part and E2 the
energy of the second part, and consider a flow of
energy (as thermal energy transfer, Q) from one part to
the other, holding the two sub-system volumes
constant. Then
dE = dE1 + dE2 = 0 (9)
dE1 = - dE2
The total entropy change, dS, arising from a flow of energy between the two parts, is

dS dS dS
dS = dE1 = 1 dE1 + 2 dE1
dE1 dE1 dE1
(E, V) dS1 dS
= dE1 2 dE1 (10)
dE1 dE2

Substitution of equation 8a and application of the second law gives

1 1
dS = dE1 0 (11)
T1 T2
The necessary condition for equilibrium is that dS = 0 for the process in which q 1 = dE 1 flows
between the two parts. This can be satisfied only if T 1 = T 2. This proves that bodies in
equilibrium have the same temperature. (This is sometimes called the zeroth law of
thermodynamics, for historical reasons, although it is a consequence of the second law.)
Suppose that T1 > T2 . Then (1/T1 - 1/T2) is negative and dE1 must also be negative, showing
that part one loses energy to part two. This proves that when two bodies in contact have
different temperatures, energy flows, as a thermal energy flow (Q), from the warmer to the cooler
body.
From equation 7 we can derive a relationship for the dependence of entropy on volume.
1 P
dS = dE + dV
T T
and therefore S P
= (12)
V E T

7/10/07 2- 39
Again consider a system constrained to constant total volume, and to have a uniform temperature.
Then, as before, we can write the total change in entropy that might result from a change in
volume of either part in the form
dS1 dS
(T, V) dS = dV1 + 2 dV1
dV1 dV1
and, because total volume is constant, dV 1 = - dV2. Substitution of equation 12 gives 5

P P
dS = 1 2 dV1 0 (13)
T1 T2
This proves that two bodies in equilibrium must have not only the same temperature but also the
same pressure. It also proves that, when two bodies in contact, at the same temperature, have
different pressures, the body at the higher pressure will tend to expand and compress the body at
the lower pressure.

ISOTHERMAL ENTROPY CHANGES. Entropy changes at a single temperature follow directly from
equation 5. (Calculations involving a temperature difference or temperature change will be
treated later.)
A change of phase, under equilibrium conditions, will be at constant temperature and
constant pressure. The thermal energy transfer, Q = Q rev, will be H. Therefore,
H
S = (14)
T

For example, the entropy change for the melting of ice was found to be 1.22 J/gK or 22.0
J/molK. The entropy of vaporization of water at 100 oC is

40 , 657 J/mol
S vap = = 108 . 96 J/mol K
373 . 15 K

More typical liquids, not involving such strong intermolecular forces, have S vap values of about
92 J/molK at their normal boiling points. This is known as Troutons rule.
The work done in a reversible, isothermal expansion of an ideal gas was found to be
(equation 10, Chapter 1) Wrev = - nRT ln V2/V1. For an ideal gas at constant temperature E = 0
and therefore Qrev = -Wrev. The entropy change is thus

5
It is assumed that T1 = T2 and thus the total entropy is unchanged by a transfer of thermal
energy between the two parts. The constant-energy restriction of equation 12 is satisfied because
any work done by one part on the other can be compensated by such a transfer of thermal energy.

7/10/07 2- 40
Qrev
(T , I.G.,rev.) S = = nR ln V2 / V1 (15)
T

If the gas expands through a pinhole or stopcock (an effusion-controlled leak) into an
evacuated container (Figure 2), no work is done because the gas exerts forces only on the
immovable walls of the container. By insulating the containers we can ensure that Q = 0. The
energy is unchanged and therefore, if the gas is ideal, the temperature remains constant. This
serves as an experimental demonstration (performed by Joule, in 1844) of the dependence of
energy only on the temperature, although it is not as sensitive as a later method by Joule and
Thomson. The entropy change of an ideal gas in such an irreversible, adiabatic 6 expansion can be
calculated as follows.
Entropy is a state function, so the entropy change depends only on the initial state and the
final state. Entropy changes are independent of the path taken between the initial and final states.
When an ideal gas expands at constant temperature from V 1 to V 2, the final state is the same
whether the process is reversible or irreversible. Therefore, the entropy change for the adiabatic
expansion of the ideal gas into a vacuum is

(I.G., T) S = nR ln V2/V1 (16)

even though Q = -W = 0 in this process.


The example just cited illustrates a general method for evaluating entropy changes in
irreversible processes. Having determined what are the initial and final states, a path is found
between those states that will be entirely reversible. Then equation 5 tells us S = Q rev/T. 7
It is of interest to calculate the entropy change for the surroundings in the reversible and
irreversible expansions above. In the reversible expansion, Q = Q rev = - Q surr = -(Q rev) surr.

6
Diabatic, like its better-known twin diabetic, implies that something passes through
(thermal energy or sugar). Thus adiabatic tells us there is no thermal energy transfer; Q = 0. In
practice we achieve adiabatic conditions by insulating the system, by maintaining system and
surroundings at the same temperature, or by carrying out the process very rapidly. Most gas
expansions are fast enough to be nearly adiabatic.
7
One sometimes sees the statement that S = Q/T for a reversible process. This is not
really wrong, but it is doubly misleading. It de-emphasizes the need to find Q rev (which is often
different from Q) and it provides no clue to finding S for an irreversible process. Equation 5, S
= Qrev /T, covers both irreversible and reversible processes.

7/10/07 2- 41
Therefore

(rev) Ssurr = - nR ln V2/V1

The adiabatic expansion into a vacuum produces no change at all in the surroundings, and
therefore

Ssurr = 0

Adding together the entropy changes for system and surroundings, we obtain, for the reversible
expansion,

(S)system + surroundings = 0

and for the irreversible expansion, into a vacuum,

(S)system + surroundings = nR ln V2 /V1 > 0

This result agrees with the requirements of the second law.


The particular irreversible path chosen is an extreme case. If the gas had been allowed to
expand against a constant external pressure (as in Figure 1, Chapter 1) the entropy change for
system plus surroundings would have been positive, but less than that for the expansion into a
vacuum. The magnitude of the total entropy change (system plus surroundings) can be taken as a
measure of the degree of thermodynamic irreversibility of any process.

INTERPRETATION OF ENTROPY. We have seen that we can measure the increase in entropy by
measuring the amount of energy transferred to the system, as thermal energy transfer, Q, divided
by the (absolute) temperature, provided we measure Q along a path between initial and final
states that is completely reversible. To know whether a process will actually occur, we must
make a similar measurement for changes in the surroundings, as we will emphasize below.
Without getting involved in the detailed calculations of statistical mechanics, it is important
to look more carefully at the meaning of entropy as a spread function. In particular, there are two
ways of measuring the spreading, or the randomness, of a state. We have seen that entropy is
appropriately called the spread function because it is a measure of the amount of spreading of
energy among the molecules of a substance. However, in addition to considering where the
packets of energy are located we should look also at where the molecules are (like the yellow and
white rice molecules we mixed earlier). Remarkably, the same measurement of Q rev /T gives
information on this second type of spreading.
Consider a sample of an ideal gas, occupying an initial volume V i (e.g., 1 m 3). Assume, quite
arbitrarily for the moment, that each molecule requires some very small volume, V o (e.g., 2 3,
most of which volume elements are unoccupied at any given instant). Then we can say there are
spaces for Vi /Vo molecules (5 x 1029 in our example). If we let the gas expand to a larger volume,
Vf , then by the same reasoning there are now spaces for Vf /Vo molecules. The important thing is

7/10/07 2- 42
that we have increased the number of available spaces by the ratio V f /V i . But we already know
that if we expand an ideal gas, the entropy increases by S = nR ln V f /V i . In other words, when
we increased the number of available, or accessible, spaces by the ratio V f /V i , the entropy
increased by the logarithm of this same ratio.
We could equally well make the same kind of argument for molecules in a crystal, where the
number of available spaces is more readily counted. If some number of stray molecules are
introduced and given an opportunity to diffuse throughout the crystal, we would obtain a
substantial increase in entropy per molecule. Or we may simply drop a crystal of salt into water.
When we return, very likely the salt molecules will have diffused throughout the water. The salt
molecules, or their constituent ions, are attracted to each other and roughly equally attracted to
water molecules, so there is negligible energy difference, but there is a very high probability of
the salt diffusing into the water, never to return to the crystal unless external conditions are
changed. It is like a drop of water on the pavement. Even though energetically the water drop
would prefer to remain with its neighbors, the lure of wide open spaces is enough to make the
drop evaporate. The process is driven by an increase in entropy, or in probability.
The crystal is highly ordered; the solution is highly disordered, or random. There are more
states accessible to the salt when it is dissolved in water, or accessible to gas molecules when a
larger volume is available to the molecules. More available, or accessible, states can be equated
to a greater probability. Thus entropy is variously described as measuring disorder, or
randomness, or probability. We can always organize molecules, forming crystals or condensing a
gas to a liquid or arranging conditions such that molecules will form living structures. To do so
always has a cost involved, specifically the cost of increased randomness in the surroundings.
The second law of thermodynamics does not tell us we cannot decrease entropy in any given
sample. It tells us only that, for the universe as a whole (system plus surroundings), entropy will
always increase.
A study of how signals may be communicated from one place to another gave rise to a new
understanding of probability under the heading of information theory. Information theory turned
out to be of major importance in understanding the meaning of entropy as well as in discovering
better ways to organize the storage and distribution of information over telephone lines, as radio
or television signals, or in computers.

Gibbs Free Energy


When investigating equilibrium or reversibility, it was necessary to find the change for the
system but also for the surroundings. Usually we prefer to ignore the surroundings most of the
time. The free energy functions provide a means of doing just that.
We began with the variables P, V, and T, to which we added energy, E, and subsequently
have added enthalpy, H, entropy, S, and now the two free energy functions, F and G, as shown in
equations 17-19 and in Figure 3.

H = E + PV (17)
F = E - TS (18)
G = H - TS (19)

7/10/07 2- 43
G = F + PV (19a)
G = E + PV - TS (19b)

We introduced the Helmholtz free energy 8 from the relationship (eqn. 4) A = F = W rev and
the observation that a change in F represents the minimum amount of work that must be done on
the system, at constant temperature, to get from the initial to the final state. For example, you
cannot put a box on a shelf without doing some work on the box (and gravitational field), nor can
you compress a gas without doing work on the gas. There is good reason for picking the correct
function. Recall that we found enthalpy, H, to be much more convenient than energy, E, when
measuring thermal energy transfers under conditions of constant pressure. For much the same
reasons, we will find G to be generally more convenient than F. The Helmholtz free energy is
more appropriate at constant volume; Gibbs free energy is better for constant pressure.

FREE ENERGY AND EQUILIBRIUM. Assume for the moment that the only work done is work of
expansion or compression, and that the composition of the system does not change. Then, from
equation 19b,
dG = dE + d(PV) - d(TS)

As we have seen, because E is a state function, dE is independent of path, so

dE = q + w = qrev + wrev

regardless of path. We may therefore substitute q rev + w rev for dE, to obtain

dG = qrev + wrev + P dV + V dP - T dS - S dT (20)

But qrev = T dS and wrev = - P dV. Therefore, quite generally,

dG = V dP - S dT (21)

We effectively split this equation into two parts, by holding constant first the temperature,
then the pressure:
G
=V (22)
P T
G
= S (23)
T P

8
For many years the Gibbs free energy and the Helmholtz free energy were not clearly
distinguished, and thus both were designated by the symbol F. In recent decades the symbol F
has been selected for Helmholtz free energy and the symbol G for Gibbs free energy. Unless
specifically indicated otherwise, the term free energy will always mean the Gibbs free energy in
this book. [Fig. 3: A becomes F.]

7/10/07 2- 44
Note, of course, that P and V naturally go together, as do
T and S.
An equally valid alternative to equation 20 is obtained
by substituting dE = q + w, giving

dG = q + w + P dV + V dP - T dS - S dT

Then, for constant temperature and pressure,

(T,P) dG = q + w + P dV - T dS
(24)

But at constant pressure we may safely assume the pressure of the system is the same as the
pressure of the surroundings (meaning the part of the surroundings that is in actual contact with,
and therefore may affect, the system), so w = - P dV, where, as usual, P is the pressure of the
system. Also, we know9 that qrev $ q. Therefore,

(T,P) dG = q - T dS = q - qrev # 0

Thus

(T,P) dG # 0 (25)

in general, and for the limiting case of the reversible, or equilibrium, process,

(T,P) dG = 0 (25a)

This shows that G must be negative, or, in the limit of a reversible process, zero. That is, G is a
minimum for an equilibrium state. This is the more convenient criterion for equilibrium that was
sought. Looking for the change in free energy, G, under our normal operating conditions of

9
We introduced the assumption that wrev # w by considering an expansion or compression
of an ideal gas, and from that and the independence of dE on path, we concluded that q rev = dE -
wrev $ q. Now we may work backward from the second law, which is one of our basic postulates,
(dS)syst + surr = qrev/T + (qrev)surr/Tsurr $ 0. This inequality must be valid for any process occurring in
the system, including if qsurr = (qrev)surr, which may therefore be assumed without loss of generality.
But we know that qsurr = - q (by combining the first law with the first-law equation), and because
we are assuming temperature is constant, we may set T surr = T. Therefore (dS)syst + surr = q rev/T -
q/T $ 0 and thus qrev $ q. It follows from this that wrev # w .

7/10/07 2- 45
constant temperature and constant pressure, we need only look at changes of the system. It is not
necessary to make separate calculations for the surroundings.
We began this section with the assumption that the only work done was P dV work, or work
of expansion or compression. Now we revisit that question. Retain the conditions of constant
temperature and constant pressure, but from equation 20 we obtain

(T, P) dG = P dV + wrev

The total work done on the system may be broken into expansion/compression work and other
forms (e.g., electrical work) which we label w. For reversible processes,

wrev = - P dV + wrev
It follows that
(T,P) dG = wrev (26)

Thus, apart from providing a convenient measure of whether the system is or is not at equilibrium
(with only P dV work), the free energy tells us how much other work can be expected from a
process. In an electrochemical cell, w is the electrical energy generated by the chemical reaction.
Work against a gravitational field or against a magnetic field would also be included in w.

IDEAL-GAS EXPANSIONS. The change in free energy of an ideal gas in an isothermal expansion or
compression is particularly important because it has been found possible to express free energy
changes of all systems, whether gases or not, in this same form. The broader theory will be
developed in the discussion of physical equilibria.
Start with the definition of G, equation 19. At constant temperature,

(T) G = H - T S (27)

In an isothermal, reversible expansion of an ideal gas, E = 0 and H = 0.

qrev = - wrev = InRT dV/V = nRT ln V2/V1


S = qrev /T = nR ln V2/V1
G = - T S = - nRT ln V2/V1
or
(T, I.G.) G = nRT ln P2 /P1 (28)
The same result could be obtained from equation 22.

G
G = dG = dP = V dP = nRT dP/ P = nRT ln P2 / P1
P T
FREE-ENERGY CHANGES IN CHEMICAL REACTIONS. A very important application of the free-
energy function is to the description of chemical reactions. Because G is a state function, free
energies can be added and subtracted in the same manner as enthalpies. It is not possible to

7/10/07 2- 46
assign meaningful absolute values to free energies, but handbook tables give values of G relative
to the elements from which compounds are formed. Thus the free energy of formation of any
element is, by definition, zero. From Table 1 we can see that the standard free energy of
formation of SO3 in its standard state (gas, at 1 atm pressure, at 25 oC) is - 370.4 kJ/mol, whereas
Gof (SO2) = - 300.2 kJ/mol. Subtracting (and setting Gof (O2 ) = 0) gives Greaction = - 70.2
kJ/mole. The negative value indicates that the reaction is spontaneous, showing that oxygen gas
will combine with sulfur dioxide to form sulfur trioxide at this temperature and pressure, at least
in the presence of a suitable catalyst.

Table 1 STANDARD ENTHALPIES AND FREE ENERGIES OF FORMATION*


Compound State Hfo Gfo Compound State Hfo Gfo
AgBr c -100.4 -96.9 H2O g -241.8 -228.6
AgCl c -127.0 -109.8 lq -285.8 -237.1
AgI c -61.8 -66.2 H2O2 g -187.8 -120.4
Ag2O c -31.1 -11.2 lq -136.3 -105.6
Al2O3 c -1675.7 -1582.3 H2S g -20.6 -33.4
Br2 g 30.9 3.1 I2 c(rh) 0 0
C diamond 1.9 2.9 NH3 g -45.9 -16.4
C graphite 0 0 aq -362.5 -236.5
CF4 g -933.6(639.9) 66.2 NH4Cl c -314.4 -202.9
CCl4 lq -139.3 -68.6 N2O g 81.6 103.2
CO g -110.5 -137.2 NO g 91.3 87.6
CO2 g -393.5 -394.4 NO2 g 33.2 51.3
CaCO3 calcite -1207.6 -1129.1 N2O4 g 11.1 99.8
aragonite -1207.8 -1128.2 NaCl c -411.2 -384.1
CaCl2 c -795.4 -748.8 aq -407.3 -393.1
aq -877.2 -816. NaOH c -425.8 -379.5
CaF2 c -1228.0 -1175.6 aq -470.1 -419.1
aq -1208 -1031.2 O3 g 142.7 163.2
CaO c -634.9 -603.3 SO2 g -296.8 -300.1
Ca(OH)2 c -985.2 -897.5 SO3 g -395.7 -371.1
aq -1002.8 -858. CH4 g -74.6 -50.5
FeO c -272.0 -244.3 C2H2 g 227.4 209.9
Fe2O3 c -824. -742.2 C2H4 g 52.4 68.4
Fe3O4 c -1118.4 -1015.4 C2H6 g -84.0 -32.0
HBr g -36.3 -53.4 CH3OH g -201.0 -162.3
aq -121.6 -104.0 lq -239.2 -166.6
HF g -273.3 -275.4 C2H5OH g -234.8 -167.9
HI g 26.5 1.7 lq -277.6 -174.8
C6H6 lq 49.1 124.5
g 82.9 129.7
*Values are in kJ/mol, for 25oC, 1 atm. The standard states for liquids (lq) and crystalline solids (c) are the pure
materials at 1 atm pressure and the standard states for gases (g) are the pure gases at 1 atm pressure in the ideal
gas state (that is, extrapolated from low pressures assuming ideal gas behavior). aq refers to the (hypothetical
ideal) 1 molal solution in water.

7/10/07 2- 47
Free energy values do not tell us anything about rates of reaction. There may be a significant
activation energy required to get a reaction to proceed, even when it is thermodynamically
allowed. (That is what catalysts are good for. By offering a different path, they may involve a
substantially lower activation energy and therefore a faster rate.) On the other hand, if G > 0,
the reaction cannot proceed, under the specified conditions, regardless of catalysts or time.

LECHATELIERS PRINCIPLE AND EQUILIBRIUM. For every chemical reaction there exists an
equilibrium point at which there is no further tendency for the reaction to proceed forward or
backward. Although the equilibrium point of most reactions is far to one side or the other, a true
equilibrium point can be calculated, and for nearly all systems the equilibrium point can be
demonstrated if the experimental techniques are sufficiently sensitive.
LeChateliers principle states that when any stress is applied to a system at equilibrium, the
system will respond in a manner that will reduce the stress. For example, the reaction

2 NO + O2!6 2 NO2

releases thermal energy (H = - 56.53 kJ/mol(NO 2)) but causes a decrease of volume, or
pressure. Therefore an increase of temperature will tend to drive the reaction to the left, so that
the reaction will absorb thermal energy, but an increase of pressure, by compression, will tend to
drive the reaction to the right to decrease the number of moles of gas and hence the pressure.
These relationships can be derived, quantitatively, from the second law.
The change in Greaction = Gproducts - Greactants with change of pressure is

G (G products Greactants )
=
P T P T

The derivative of the sum is the sum of the derivatives:

G Gproducts Greactants
=
P T P T P T

From equation 22, these derivatives are the respective volumes:

G
= V products Vreactants
P T

or

G
= V (29)
P T

7/10/07 2- 48
Similarly, from equation 23,

G
= S (30)
T P

This equation tells how G (measured at some constant temperature and pressure) depends upon
the temperature at which the process occurs. The S on the right-hand side is similarly to be
measured at a constant pressure and at a constant temperature (the temperature at which the
derivative, or the slope of G vs. T, is determined). We may therefore substitute for - S the
quantity (G - H)/T.
G G H
= (30a)
T P T

To find the shift of equilibrium point we set G = 0 and obtain

G H
= (31)
T P T

The first result, equation 29, says that if there is a volume increase during the process, a pressure
increase will cause an increase in G, making the process less spontaneous. The second equation
(31) says that if the process is endothermic (H positive), a temperature increase will make G
more negative, thus making the process more spontaneous. The equations apply to phase
changes or other equilibrium processes as well as to chemical reactions. 10 The importance of the
equations is that they give a quantitative description of how G changes, and therefore of how
the equilibrium point changes.

Free Energy and the Energy-Entropy Battle


For the formation of ammonia, we may write the equation

N2 +3/2 H2&6 NH3

G = - 16.4 kJ/mol, H = - 45.9 kJ/mol, and S = -99.2 J/molK. The reaction clearly is
spontaneous at room temperature and pressure, but it doesnt go. The activation barrier is too
great. From LeChateliers rule, we can see that an increase in temperature, to overcome the

10
It is assumed that there are no other changes that might cause a reverse effect. For
example, changes in the relative concentrations of the reactants, or addition of an inert gas, are
not within the scope of equations 29 and 30 or 31.

7/10/07 2- 49
activation barrier and reaction kinetics, will make the reaction less spontaneous, but an increase in
pressure will drive the reaction back to the right. The synthesis of ammonia is therefore carried
out at high temperature (to get the reaction to go) and at high pressure (to recover yield).
We can look at the reaction in a somewhat different way. Write
G = H - TS = -45.9 kJ/mol - T x (-99.05 J/molK)
Examining the numbers suggests that the reaction as written (25 oC and 1 atm of each gas) is
spontaneous because the reactants give off thermal energy (H = Q < 0 at const. P) without a
major change in the thermal energy of the system (T is constant). The system, as a whole, is
falling into a potential well, giving off energy that is conducted to the surroundings as thermal
energy. On the other hand, this drop in internal energy drives the system to a more ordered
internal arrangement two gas molecules collapse into one gas molecule, something like a
crystallization (loss of internal degrees of freedom, or a more orderly arrangement). Thus if we
raise the temperature (neglecting for the moment the change in values of H and S), the extra
degrees of freedom of the molecules of the elements become more important (for example, at
constant pressure they will occupy a larger volume and as speeds increase, the spread in speeds
will increase as well), so the condensation of elements to compound(s) becomes less probable.
They must be driven together by increasing the pressure.
Of course, common sense tells us much the same thing, qualitatively. At high temperatures,
compounds tend to decompose. At low temperatures, substances tend to condense, to liquids or
to crystals. Energy differences are important at low temperatures. The spread function, entropy,
is increasingly important as the temperature is raised.
A rough rule of thumb is that formation of 1 mole of gas molecules raises the entropy by
about 125 J/molK (compare Troutons rule, p. 33, which is for vaporization only at the normal
boiling point). At room temperature, the resultant increase in TS would be on the order of 40
kJ/mol of gas formed. This is a small term compared to the enthalpy change of very exothermic
reactions, such as most combustions, but it is not small for many other reactions.
Entropies of reaction for a few representative chemical reactions are given in Table 2.
Table 2. STANDARD ENTROPIES OF REACTION, 25oC
Reaction So, J/molK
2 Al (c) + O2 ----- Al2O3 (c) -313.3
C (gr) + O2 ----- CO2 +3.0
CO + O2 ----- CO2 -86.36
H2 + Cl2 ----- 2 HCl +19.6
H2 + F2 ----- 2 HF +14.1
H2 + I2 (c) ----- 2 HI +166.2
H2 + O2 ----- H2O (lq) -163.1
H2 + S (rh) ----- H2S +43.14
N2 + 3H2 ----- 2 NH3 -198.2
S (rh) + O2 ----- SO2 +11.2
3 C2H2 ----- C6H6 (lq) -333.3

Entropy Changes with Change of Temperature

7/10/07 2- 50
The fundamental equation for entropy change, equation 5, gives the infinitesimal entropy
increase for the absorption of an infinitesimal amount of thermal energy at some temperature T.
If the temperature changes continuously during the process, the sum of the entropy changes can
be written as the integral

q rev
S = dS = T
(32)

It is not at all obvious from equation 32 how this integral is to be evaluated, but it can be handled

quite easily as follows:


qrev q dT q dT dT
S = = rev = rev = C (33)
T T dT dT T T
The heat capacity, C, is often nearly independent of temperature (over temperature intervals of
tens of degrees) so the equation can be simplified to the form
T2
S = C ln (34)
T1

Note that the heat capacity has not been given a subscript in equations 33 or 34; this is because
no conditions such as constant pressure or constant volume have thus far been specified. If
pressure is constant, CP is inserted in these equations; if volume is constant, CV is required. If
neither pressure nor volume is constant, it may be necessary to find an alternative path consisting
of constant pressure and constant temperature segments, or constant volume and constant
temperature segments. It may be, too, that it will be simpler to return to equation 32. For
example, in a reversible, adiabatic expansion or compression, q = q rev = 0, and therefore it is clear
that the integral, and the entropy change, must be identically zero. To apply equation 34 one
would have to define an adiabatic heat capacity that would always be zero.

7/10/07 2- 51
The entropy change for the transfer of thermal energy between two bodies at different
temperatures requires a different approach. In Figure 4, the body at the left is at a temperature
T1, the body at the right is at a lower temperature T2 , and each body is assumed large enough that
its temperature will not be appreciably affected by the flow of thermal energy.
The primary method of energy transfer in such a situation (e.g., between the Sun and the
Earth) is by radiation. But the radiation process is not at equilibrium. (Radiation at equilibrium is
called black body radiation and requires the radiation field to be in equilibrium with the radiating
body, which cannot be the case if radiation is being leaked to the low-temperature body.) We
must therefore look for a reasonable substitute that will give us an answer suitably close to reality.
We achieve this approximation to reality by a thought experiment. Add a thermal insulator
connecting the two bodies. When a steady state has been achieved, the temperature will vary
continuously with the distance, as shown in the lower part of Figure 4, but the temperature will be
constant at each point in the system even though a small amount of thermal energy is flowing past
each point. The entropy change is the sum of three terms: the entropy change of the left-hand
body at T1; the entropy change for the insulator; and the entropy change for the right-hand body
at temperature T2.
S = S1 + Sinsulator + S2

For an amount of thermal energy transferred, Q, the first term is

Q
S1 =
T1

where S1 is the entropy change of the body at T1 . Similarly, the entropy change for the body at
temperature T2 is
Q
S 2 =
T2

The insulator may be considered part of the surroundings and need not be further included in the
calculation for the system.11 Then for the system, which consists of the two bodies at the two
temperatures,
Q Q 1 1
S = S1 + S 2 = + = Q > 0 (35)
T1 T2 T2 T2

Superficially the thermal insulator added to the problem of Figure 4 plays no significant part
in the calculation. It can make little difference to the state, or the entropy, of a body that loses a
given amount of thermal energy at the temperature T 1 what happens to this energy after it is gone.

11
Actually, the entropy change of the insulator is zero, after the steady state has been
reached, because the state of the insulator is not changing.

7/10/07 2- 52
Similarly, to the body absorbing a certain amount of thermal energy at the temperature T 2, it can
make little if any difference to the state, or entropy, of the body, where the energy came from. At
the same time, we must recognize we have introduced an approximation by assuming the energy
transfer was thermal energy transfer under equilibrium conditions. If, for example, the energy is
transferred by radiation, higher-temperature radiation has a greater chance of knocking electrons
free as photoelectrons, so the state of the receptor may depend, at least to some extent, on the
temperature of the higher-temperature radiator. The closer the experimental situation approaches
true equilibrium, the more assurance we can have that the answers are adequate. For example, in
the paragraph preceding equation 6, in which thermal energy is transferred to the system (ice)
from the surroundings (a thermostat) at a slightly higher temperature, there is no reason to doubt
the conclusions.

Problems
1. Figure 5 shows a capillary tube dipping into a dish of water. The tubing is bent, and water
dripping from the suspended open end turns a paddle wheel before returning to the dish. By
covering the apparatus, water can be retained and the energy required is returned to the dish by
thermal conduction from the surroundings. What is wrong?

2. Figure 6 shows a wheel on which are mounted small bar


magnets. The force between the fixed magnet and the
individual magnets attached to the wheel may be taken as
proportional to 1/d, where d is the distance separating the
magnets. Show that the force between the fixed magnet
and the nearest magnet is smaller than the sum of the forces
pulling the magnets along the right-hand side. That is, 1/d 1
< 1/d2 + 1/d3 + . A magnetic shield (readily available) is
placed as shown. The wheel then rotates
counterclockwise. What is wrong?

7/10/07 2- 53
3. Find S when 2 mol of an ideal gas at 25oC is expanded into vacuum to 10 times its original
volume.
4. A vessel containing 5 mol of an ideal gas, A, is connected to an identical vessel containing 5
mol of another ideal gas, B, and the two are allowed to reach equilibrium. For this process, at
27oC, find
a. S for the gas A.
b. S for the gas B.
c. S for the entire system (S(A) + S(B)).
d. What would S be if A and B were the same gas?
5. If a gas at pressure P and volume V1 is mixed isothermally with another gas at pressure P and
volume V2 to give a mixture at pressure P and volume V = V1 + V2 , the mole fractions are N1 =
V1/(V1 + V2) and N2 = V2/(V1 + V2 ). Show that the entropy of mixing of ideal gases in this
fashion is

S = nR i
N i ln N i

where n = Gni is the total number of moles and Ni = ni /n is the mole fraction of the i th gas.
6. A vessel containing 5 mol of an ideal gas, A, is connected to an identical vessel containing 10
mol of an ideal gas, B, and the two are allowed to reach equilibrium. For the process, at 27 oC,
find
a. S, assuming A and B are different gases.
b. S, assuming A and B are the same. Explain the distinction between these two cases and
between these and problem 4.
7. Which of the following have the higher value of S?
a. CO2 at 25oC, 1 atm or dry ice at 1 atm
b. a coiled spring or the spring relaxed
c. 1 g of liquid water at 25oC or 1 g of water vapor, at the vapor pressure of water at
25oC.
d. silica glass or quartz (crystalline silica)
8. When a gas escapes through a pinhole into an evacuated chamber, the first gas to escape is
later compressed by the following gas. Explain why W = 0, despite this compression.
9. Calculate G for the processes described in problems 3, 4, and 6.
10. The heat of vaporization of cyclohexane, C6H12, is 30.46 kJ/mol at its normal boiling point,
80.7oC.
a. Calculate S for the vaporization of 1 mol of cyclohexane at 80.7 oC and 1 atm. Compare
with the Trouton constant.
b. Calculate G for the same process.
11. Calculate the free energy change when 3 mol of Ar is compressed isothermally from 2 to 6
atm at 50oC.
12. The standard free energies of formation of ethylene (C 2H 4) and of ethane (C2H6) at 25 oC and
1 atm are 68.4 and -32.0 kJ/mol. The standard enthalpies of formation are 52.4 and -84.0
kJ/mol. For the reduction of ethylene with H2 to give ethane,

7/10/07 2- 54
a. find Go
b. find So.
13. Calculate So, the entropy change under standard conditions at 25o C, for the reaction of
carbon (graphite) with fluorine gas to give the gas CF 4, using values as necessary from Table 1.
14. For each of the following reactions at 25oC calculate (using Table 1)
a. Ho b. Go c. So

4 FeO + O2 ----6 2 Fe2O3


2 Fe3O4 + O2 ----6 3 Fe2O3
3 FeO + O2 ----6 Fe3O4
Which form of iron oxide FeO, Fe3O4, or Fe2O3 is most stable in the presence of an oxygen
atmosphere at this temperature? Explain.
15. Would Fe3O4 (see problem 14) become more or less stable, with respect to FeO and oxygen,
as temperature is increased? Explain.
16. For the reaction at 25oC, with each gas at 1 atm,

3 H2 + N2 ----6 2 NH3

a. Find H. b. Find E.
c. Find G. d. Find S.
e. Is the reaction as written, at constant T and P, spontaneous?
f. Does the reaction as written take up or give off thermal energy?
g. Find the pressure, P, of NH3 for which G = 0.
17. In the following reactions, would you expect S to be positive or negative? Explain.
a. Ag2O (s) ----6 2 Ag (s) + O2 (g)
b. 2 CO (g) + O2 (g) ----6 2 CO2 (g)
c. 2 C (s) + O2 (g) ----6 2 CO (g)
d. CaH2 (s) + 2 H2O (l) ----6 Ca(OH)2 (s) + 2 H2 (g)
e. H2 (g) + I2 (g) ----6 2 HI (g)
f. 2 H2 (g) + O2 (g) ----6 2 H2 O (l)
g. n C2H4 (g) ----6 (C2H4)n (polyethylene)
18. A Dewar flask contains 500 ml of water at 25oC. To the flask is added 100 g of ice at - 6o C.
For the process which then occurs, find
a. H
b. S
c. the final temperature
19. Find
a. S b. H c. E
for the process of subliming 6 g of ice at - 10 oC and warming the vapor to 200oC and 4 atm.
20. One kilowatt hour of energy passes from a heater, at a temperature of 1000 K, to a room at
27oC. Find S.
21. For the freezing of 5 g of supercooled water, at - 10 oC, to ice at - 10oC, calculate
a. H b. S c. G

7/10/07 2- 55
22. The standard free energies (1 atm) of NO 2 and N2O4 at 25 oC are 51.3 and 99.8 kJ/mol For
the reaction
2 NO2 (1 atm) ----- N2 O4 (P atm)
a. find G if P = 1.
b. find the value of P (i.e., the pressure of N 2O 4) that would make G = 0.
23. In a reversible, adiabatic expansion, q = 0. Show, by setting dE = n C V dT and w = - P dV
that, if P = nRT/V, the equation w = dE can be integrated (after dividing through by T to separate
variables) to give
T2 V
(I.G., rev, adiabatic) CV ln = R ln 1 (36)
T1 V2

24. Show that, by replacing dV by d(nRT/P), the equation for a reversible, adiabatic expansion
integrates (after separating variables) to give

T2 P
(I.G., rev, adiabatic) C P ln = R ln 2 (37)
T1 P1

(Note that alternatively equation 37 may be obtained from equation 36 by replacing V 2 by nRT 2
/P2 and V1 by nRT1 /P1.)
25. Show that equations 36 and 37 for an adiabatic, reversible expansion of an ideal gas can be
put into the following forms:
a. TV R/CV = constant, or T1/T2 = (V2/V1)R/C V.
b. PV = constant, or P1/P2 = (V2/V1) , where = CP /CV .
26. An increase of entropy may correspond to an increase of randomness of velocities, or
momenta, rather than of positions. For example, when a gas is heated at constant volume there
can be no change in the spatial disorder but there is more of a spread in the molecular speeds.
What conclusions can be drawn concerning the change in temperature when an isolated
supersaturated solution spontaneously separates into two phases by precipitation of solute?

7/10/07 2- 56
3 Physical Equilibria
The power of thermodynamics lies in its applicability to all substances in all states. It is not
limited to such abstractions as ideal gases, ideal solutions, or perfect crystals, although certain of
the equations take on especially simple forms for such special cases. In the following discussions
of physical equilibria and chemical equilibria the exact thermodynamic equations will be derived
wherever possible. This will enable a rational choice of approximations for each application and a
clearer understanding of when, and what, approximations are being introduced.

General Conditions for Equilibrium


The fundamental requirement for a system to be at equilibrium, with respect to a certain
process, is that under the existing conditions the process should be thermodynamically reversible.
The requirement is, from the second law, that the entropy of the system and the surroundings
should remain unchanged for small changes in the state of the system. This is a basic statement,
from which other conditions can be derived. It is not necessarily, however, the most useful way
of stating the condition of equilibrium. In this chapter the conditions for equilibrium not involving
chemical change will be put into a variety of forms. Depending upon the process considered, one
or another of these will be found most convenient.

CONDITIONS OF TEMPERATURE, PRESSURE, AND FREE ENERGY. It was shown in the previous
chapter that the second law requires that two bodies in equilibrium should have the same
temperature and the same pressure. It will therefore be assumed in the following discussion that
these conditions are satisfied. We will be interested in the question of what that temperature and
pressure must be, and how these parameters are affected by each other and by changes in
chemical composition.
It was also shown that a necessary and sufficient condition for equilibrium at constant
temperature and pressure is that the free energy of the system be a minimum, and hence that the
free-energy change of the system be zero for small changes in the system. The condition will
form the basis of the derivations in this chapter.

CLAPEYRON EQUATION. Pure water freezes at 0oC and boils at 100oC, but only under normal
conditions of 1 atm pressure. An increase of pressure will lower the freezing point but will raise
the boiling point. Both effects are easily predicted by the Clapeyron equation, which is to be
developed.
Let some substance exist in two phases, A and B, in equilibrium at some temperature and
pressure. This might be the equilibrium between ice and liquid water, or liquid water and its
vapor, or some other solid-solid, solid-liquid, solid-vapor, or liquid-vapor equilibrium. Then the
condition for equilibrium is

G = GB - GA = 0 (1)

7/10/07 3- 57
We change the temperature and simultaneously change the pressure in such a way that
equilibrium is maintained. This requires that G A and G B change by the same amount.

dGA = dGB (2)

From equations 21, 22, and 23 in Chapter 2, the change in free energy with temperature and
pressure is

dGB = VB dP- SB dT
dGA = VA dP- SA dT
and therefore
VB dP - SB dT = VA dP - SA dT

This can be rearranged to give

dP SB SA S
(Equil. between two phases)
= = (3)
dT VB VA V

Because equilibrium is maintained, G = 0 and S = H/T. This gives the Clapeyron equation in
its usual form:
dP H
(Equil. between two phases) = (4)
dT TV

From this equation the change in a freezing point, or other phase transition, with change of
applied pressure can be calculated.
For example, the freezing point of water under 1 atm of air is exactly 0 oC. The freezing point
under 20 atm pressure can be calculated as follows. 1

P 19 atm H - 334 J/g 1 atm 1 Pa


= = = x
T T TV 273 K x 9.035 x 10 m / g 1 .013 x 10 Pa 1 J/m 3
-8 3 5

19 x 273 x 0.09035 x 10 -6 x 1.013 x 10 5


T = K = 0.14 o C
334
where we have inserted the value 0.917 g/cm 3 for the density of ice and 1 atm = 1.01325 x 105 Pa
(1 Pa = 1 pascal = 1 N/m2 = 1 J/m3).

1
The substitution of P/T for dP/dT is justified because the change in temperature is
small. More generally, one could carry out the integration to obtain IdP = P = H/V IdT/T =
H/V ln T2/T1. However, ln T2/T1 = ln (1 + (T2 - T1 )/T1 ) and for small values of x, ln(1 + x) = x.
Therefore, for (T2 - T1)/T1 small, P = H/V T/T1 or P/T = H/TV.

7/10/07 3- 58
Escaping Tendency
The thermodynamic picture of equilibrium is static, with no tendency for change in the total
amounts of material in the two phases. The molecular picture of equilibrium is dynamic, with
molecules continually passing back and forth from one phase to the other; but this dynamic model
is entirely consistent with the static concept of equilibrium because the number of molecules
passing in one direction is just equal, at equilibrium, to the number of molecules passing in the
other direction.
Employing either the macroscopic and static picture or the dynamic molecular view, we may
say that at equilibrium the tendency for molecules, or material, to escape from one phase to the
other is the same as the tendency for molecules, or material, to escape in the reverse direction.
In this language, the condition for equilibrium is that the escaping tendency for any substance
must be the same in all phases in equilibrium.
The free energy is a measure of this escaping tendency. The higher the free energy, the
greater the escaping tendency and the lower the stability of a phase. The minimum free energy, or
minimum escaping tendency, represents the most stable, or equilibrium, condition. However, for
many problems the free energy itself is not convenient. First, we cannot know an absolute value
for the free energy. We know only values relative to arbitrary standard states. Second, whatever
the arbitrary value assigned for a standard state, the value of G will become negative and infinite
as the pressure of a gas or concentration of a solute approach zero. Gases at low pressures, and
dilute solutions, are too important to permit such awkward behavior of the function chosen to
describe them. We turn, therefore, for this purpose to the vapor pressure or to an idealization of
the vapor pressure.

VAPOR PRESSURE AND CLAUSIUS-CLAPEYRON EQUATION. At room temperature (25oC) water


molecules leave the surface of the liquid at a rate equal to that at which molecules of the vapor
strike the surface, if the vapor is also at room temperature and at a pressure of 23.756 torr 2; that
is, at 100% relative humidity. If the temperature is raised the molecules in the liquid have higher
speeds and more of them will escape from the surface. The molecules in the vapor will also have
higher speeds and they will strike the surface more frequently. Is this enough to maintain
equilibrium? It is not, because as so often happens, we encounter an exponential dependence
relating energy and temperature.
The number of molecules escaping from the surface depends on the number having speeds
above some limiting value. The number of molecules having speeds above a limiting energy value
depends exponentially on the temperature, whereas the increase in average speed in the vapor
depends only on the square root of temperature. Therefore, in order to maintain equilibrium, the
density of molecules, or the pressure, must be increased, and in fact it must increase
exponentially.
An approximate equation describing the change of vapor pressure with temperature can be
derived from the Clapeyron equation by assuming the vapor to be an ideal gas. The volume of

2
1 torr = 1 mm of Hg = 1/760 atm = 133.3 Pa

7/10/07 3- 59
one mole of the vapor is RT/P and the volume of the liquid is negligible in comparison, so RT/P
may be taken as V. Thus
dP H P
=
dT T RT
or
dP d ln P H
(Ideal vapor in equil. with liquid or solid) = = (5)
PdT dT RT 2

This is called the Clausius-Clapeyron equation. It serves for calculations of changes of boiling
point, or sublimation point, with change of pressure, which is equivalent to saying that it gives the
variation of the vapor pressure of a solid or liquid with change in temperature. Vapors at their
condensation points are not ideal gases, so it is only an approximation, but it is often adequate.
The percentage change in vapor pressure, P/P, for a one-degree change in temperature for
water at its normal boiling point, can be found from the Clausius-Clapeyron equation. The heat
of vaporization of water is 2.258 kJ/g or 40,657 J/mol, at 100 oC. Therefore,

P 40,657 J/mol x 1 K
= = 0.035 = 3.5%
P 8.314 J/mol K x (373) 2 K 2
For larger ranges of temperature the differential form of the Clausius-Clapeyron equation
(equation 5) must be integrated to give the equation

P2 HT
ln = (6)
P1 RT1T2

This form assumes that the heat of vaporization (or heat of sublimation) is a constant over the
temperature range T = T2 - T1. It should be recognized that H does depend on temperature,
because the heat capacities of condensed phase and gas phase are different. The change in H is,
however, usually small compared to H itself if T is small, and the error can be further reduced
by substitution of an average value for H into equation 6.
From equation 6 the vapor pressure of water at room temperature would be calculated as
follows:
P2 2258 x 18 x(-75)
ln = = 3.30 = ln 0.0369
760 8.314 x 298 x 373
P2 = 760 x 0.0369 = 28.1 torr

Substitution of an average value for H of about 2.34 kJ/g would change the final answer to P 2 =
25.2 torr. The experimental value is 23.756 torr. Over a smaller temperature range better
agreement would be expected.
The pressure appearing in the Clausius-Clapeyron equation is the pressure of the (ideal) pure
vapor in equilibrium with the liquid. It is not necessarily the same as the total pressure applied to

7/10/07 3- 60
the liquid. For example, the pressure on water in the laboratory is 1 atm, applied by the air, but
the pressure of the water vapor at equilibrium is only 23.756 torr, or about 0.03 atm. The
pressure of a single component is often called the partial pressure, to distinguish it from the
total pressure. It is the equilibrium partial pressure, or vapor pressure, that is a measure of the
escaping tendency of the liquid. The escaping tendency of the liquid is affected, but only slightly,
by the total applied pressure.

FUGACITY. When the vapor is ideal there is a simple, exact relationship between vapor pressure
and the thermodynamic properties, especially the free energy, G (equation 28, Chapter 2). It
would be a great help to have an exact relationship for real gases and the escaping tendencies of
liquids and solids that could be employed for thermodynamic calculations.
The logical requirement for this new function, f, is that it should coincide with the vapor
pressure of a solid or liquid, and with the partial pressure of a vapor, whenever the vapor is ideal,
but it should retain the same functional relationship to G even when the vapor is non-ideal. This
new measure of escaping tendency may be considered a more practical vapor pressure. We define
the function, f, by the two equations

f2
G = RT ln (7)
f1

and

f
Lim =1 (8)
P 0 P
An alternative form of the first of these equations is to write

G = RT ln f + B(T) (7a)

with the unknown function B(T), depending on temperature, added so that absolute values can be
assigned to f even though absolute values are not known for G. The function f is called fugacity,
from the same Latin root (fugere) as fugitive. The value to be inserted for G when the
substance is not pure is an effective value, to be discussed later.
From the defining equations (7 and 8) a value for the fugacity can be found experimentally
for any substance. At a sufficiently low pressure of the vapor, fugacity will equal pressure. The
change in free energy with pressure can be found from equation 22, Chapter 2,

G = IV dP (9)

If the volume is known for each pressure, this integral can be evaluated (by numerical integration
if necessary) and the fugacity at the higher pressure can be found from the change in free energy:

7/10/07 3- 61
P2 f2 f
G = VdP = RT ln = RT ln 2 (10)
P1 f1 P1

In this equation, P1 is chosen to be sufficiently small so that P1 = f1 . Alternatively, the volume


may be written
RT
V = +
P

where is a correction term that varies with pressure and is zero for zero pressure. Then, for any
P1 and P2,

P2 P2
G = RT ln + dP (10a)
P1 P1

In practice these somewhat tedious integrations are seldom necessary.


Fugacity tables are available for gases. When the actual volume of a gas is known at a
certain temperature and pressure, the fugacity may be approximated by means of the equation

f . P2V/RT

or, letting Pi = RT/V,


f P
(11)
P Pi

This equation can be derived from van der Waals equation or other similar equations of state by
neglecting second-order correction terms. Note that the fugacity differs from the true pressure of
a non-ideal gas in the opposite direction from that of the ideal pressure, P i = RT/V.
In many applications we need not know the absolute value of the fugacity but only a relative
value compared to some standard state. The ratio, f /f o, where f o is the fugacity in the particular
standard state chosen, is called the activity. The activity is a dimensionless ratio; the value of the
ratio depends on the choice of standard state as well as the state of the substance being described.
Different standard states are frequently employed in different problems, or often even in the same
problem, so different values of the activity are obtained for a given state of any substance.
In the discussions that follow the reader may, without appreciable error, substitute the term
vapor pressure for fugacity, or else the term partial pressure if the substance is itself a vapor.
The equations are exactly true as given and are approximately true if vapor pressures (or partial
pressures) are chosen as approximations to the fugacity. (Note, however, that the total applied
pressure may be very different from the vapor pressure, or fugacity, of a solid or liquid.)

EXACT CLAUSIUS-CLAPEYRON EQUATION. The Clapeyron equation (equation 4) is an exact

7/10/07 3- 62
expression relating total applied pressure to the temperature. From it an approximate equation
(equation 5), the Clausius-Clapeyron equation, was derived by assuming the vapor to be ideal and
by neglecting the volume of the condensed phase. It is now possible to derive an exact form of
the Clausius-Clapeyron equation, which will prove invaluable in later applications where the
assumptions of ideal vapors will be unnecessary or inappropriate.
Recall that fugacity is a measure of tendency to escape. Now we wish to consider the process
of escape of molecules from a condensed phase into the vapor phase at a sufficiently low pressure
(of vapor) that interactions in the vapor are negligible. Thus we take as the escaped state a
pressure, P*, sufficiently low that f* = P*. If H* is the enthalpy of the vapor at such a low
pressure, and H is the enthalpy of the substance (at the equilibrium pressure) in the condensed
phase (liquid or solid), the Clausius-Clapeyron equation may be written in the exact form
d ln f H * H
(Vapor in equil. with liquid or solid) = (12)
dT RT 2

The enthalpy difference, H* - H, is the heat of vaporization (or the heat of sublimation) plus a
small correction term for the change of pressure from the equilibrium pressure to the very low
pressure.
P* H
H * H = H vap + dP (13)
P
P T

The second term, the variation of enthalpy of the vapor with pressure, is sufficiently small that it
can usually be neglected; for an ideal gas it would be identically zero. Also, for an ideal gas the
fugacity is equal to the vapor pressure, so for an ideal gas, equation 12 becomes identical with
equation 5.
The exact Clausius-Clapeyron equation may be derived as follows. The free-energy change
in going from the condensed phase, at some pressure P, to the vapor at the very low pressure P*
is
f*
G = G * G = RT ln
f

(For the vaporization step there is no change in free energy, or fugacity, but there is a change in
free energy with change of pressure.) This is easily rearranged to give
G G*
ln f = + ln f *
RT

To find the temperature dependence we take the derivative with respect to temperature (with
pressure constant and maintaining equilibrium between the two phases):

7/10/07 3- 63
d 1 1 d
= (G G *) (G G *) + d ln f *
d ln f
+
dT dT RT RT dT dT
The last term is zero, because f* = P* was chosen as an arbitrary but fixed value, independent of
temperature. Employing equation 23, Chapter 2, the remaining terms give

d ln f G * G S * S
= +
dT RT 2 RT
H * H
=
RT 2

which is the equation (12) we sought.

Colligative Properties; Laws of Dilute Solutions


The liquid phase is the most difficult to understand and to describe quantitatively. The
molecules cannot be considered independent of each other, as in a gas, nor are they arranged in a
regular, fixed structure, as in a crystal. There is great individuality of behavior of pure liquids,
and of solutions, so that few generalizations can be made.
The one area in which thermodynamics has been most extensively and successfully applied to
solutions is that of very dilute solutions. The laws that can be derived for very dilute solutions are
largely independent of the nature of the solute molecules but depend on the number of such solute
molecules. The properties are called colligative (meaning collective). The dilute-solution laws
are commonly applied to solutions of moderate or even high concentrations, much as the ideal-
gas law is applied to real gases as an approximation. Indeed, just as the ideal-gas laws were
suggested by measurements on real gases, the dilute-solution laws were suggested by
measurements on non-dilute solutions. However, the errors encountered in extrapolation of the
equations to moderate or high concentrations can be quite large. It is best to remember that these
equations must hold for infinitely dilute solutions and may hold for greater concentrations.
In the following derivations and formulas it will be convenient to call the major constituent
the solvent and designate it by the subscript 1; minor constituents will be called solutes and
designated by subscripts 2, 3, etc. Concentrations will be expressed as mole fractions or, later, as
molarity or molality. The mole fraction of any component in a mixture is the number of moles of
that component divided by the sum of the number of moles of all components. Letting n 1, n 2,
represent the number of moles of components 1, 2, etc., the mole fraction, N i, of the i th
component is
ni
Ni =
n1 + n2 +
Mole fractions are also sometimes represented by the symbol X i. Molarity is the concentration
expressed as moles of solute per liter (or 1000 cm 3) of solution. For example, 0.3 mol of NaCl in

7/10/07 3- 64
sufficient water to give one liter of solution gives a 0.3 molar solution, written 0.3 M. Molality
is the concentration expressed as moles of solute per kilogram of solvent. 3 Thus 0.3 mol of NaCl
in 1 kg of H2O gives a 0.3 molal solution, written 0.3 m. For dilute water solutions, molarity
and molality are nearly the same, but for concentrated water solutions or, especially, solutions in
other solvents, the molarity and molality may be very different.

DILUTE-SOLUTION POSTULATE. The laws of dilute solutions rely on a single postulate, which has
been demonstrated to have wide validity for solutes that do not dissociate upon dilution. Thus
the following equations apply to sugar or to Na+ ions and Cl- ions (because NaCl is fully
dissociated at moderate concentrations) in water, or to solutions of polymeric materials, such as
polystyrene, even though these may have molar masses of hundreds of thousands. The range of
validity that is, what constitutes a very dilute solution will differ from one solute to another,
so some judgment is required is estimating the accuracy of the dilute-solution laws in particular
examples. The methods of statistical mechanics show that the condition to be satisfied is that
solute-solute interactions should be negligible as compared with solute-solvent and solvent-
solvent interactions. When the solute does dissociate upon dilution, as does acetic acid, the
dilute-solution laws generally can be applied to the dissociation products. In practice it is often
adequate to consider the solute to be a mixture of undissociated and dissociated solute and treat
these as different solutes.
If the fugacity of a solute is plotted against concentration it is obvious that the fugacity (or
vapor pressure) of the solute must decrease to zero as the concentration of the solute decreases to
zero. There are, in principle, three possibilities, as shown in Figure 1. The slope at the origin
might be infinite, or zero, or have some finite, non-zero value. On the basis of many experimental

3
By convention, variables (such as mole fraction, or mass) are written in italics, whereas
units are written in roman type faces (e.g., g and mg for gram and milligram). Concentrations
(which are variables) are an exception in that the units (molarity or molality) are written in italics.

7/10/07 3- 65
measurements, the following statement is put forth as a postulate: For all solutes that do not
dissociate upon dilution, the slope of fugacity against concentration, df/dN 2, at the origin is
finite and non-zero.

HENRYS LAW. In 1803 William Henry proposed, on the basis of his measurements, that the
solubility of a gas in a liquid increases in proportion to the pressure of the gas. This result can
be easily derived from the dilute-solution postulate, which can be written

df2
=k (14)
dN2 N2 0

The number k is the slope at the origin, which will depend on the solute and on the solvent, and
the notation N2 !6 0 indicates that the solution is very dilute. (That is, as N2 approaches zero the
equation must become valid.) A derivative is, by definition, the limiting value for the ratio of the
changes in the dependent and independent variable or, in this case, the limiting value, as N 2 is
made small, of the ratio f2/N2. But f2 cn be written f2 - 0, for a small increment of fugacity at
the origin, and N2 can similarly be written N2 - 0. With these substitutions, we write

df 2 f f 0 f
= 2 = 2 = 2
dN2 N2 0 N 2 N2 0 N 2 0 N2 0 N 2 N2 0

Combining this result with equation 14, we obtain

( N 2 0) f2
=k
N2

or
(N2 0) f 2 = kN2 (15)

For a very dilute solution the fugacity, and vapor pressure, of the solute are very small and
therefore the fugacity is equal to the vapor pressure.

( N 2 0) P2 = kN 2 (15a)

This is Henrys law. For dilute solutions the concentration, c 2, expressed as molarity or molality,
is proportional to the mole fraction, N 2, so Henrys law may be written in the more general form

(c2 0) f 2 = kc2 (15b)

or

7/10/07 3- 66
(c2 0) P2 = kc2 (15c)

The value of k depends on the solute, the solvent, the temperature, the units in which c 2 is
expressed (mole fraction, molarity, or molality), and the units in which f 2 or P 2 is expressed
(usually atm, Pa, or torr). Small deviations from equation 15c may be observed at moderate
concentrations because the vapor is not ideal; large deviations may occur because the solution
may not obey equation 15b (or 15) in solutions that are not highly dilute.
An example of Henrys law is provided by the human respiratory-circulatory system. Blood
entering the lungs is exposed to air containing approximately 0.20 atm oxygen, and it becomes
saturated, at this pressure, with oxygen. When the blood reaches the capillaries, the pressure of
oxygen in the surrounding tissues is less than 0.20 atm, so oxygen is given up by the blood to the
surrounding tissues. Meanwhile, the blood picks up carbon dioxide at a comparatively high
pressure in the tissues surrounding the capillaries and loses carbon dioxide in the lungs where the
partial pressure of the CO2 is comparatively low. Both gases are bound chemically within the
blood (with hemoglobin or as carbonates), but that equilibrium is controlled by the fugacity of the
free gas in the liquid phase, which is controlled in turn, through Henrys law, by the partial
pressure of the gas in the surrounding medium (the air of the lungs or the fluids around the
capillaries.).

NERNSTS DISTRIBUTION LAW. When two solutions, containing the same solute but different,
immiscible solvents, are brought to equilibrium, the final concentration of the solute will generally
be higher in one solvent than in the other. However, the ratio of the concentrations remains
unchanged if more solute is added, provided the solutions are dilute.
Consider one solute distributed between two immiscible solvents, A and B (Figure 2). The
concentration of the solute in solvent A is c 2A and the concentration of the same solute in solvent
B is c2B. From Henrys law,

(c
A
2 0) f2A = kAc2A
(c
B
2 0) f2B = kBc2B
In order that the system be at equilibrium, the free energy of the solute must be the same in the
two solutions. This requires that the fugacity of the solute be the same,

f 2A = f 2B
and therefore

k Ac2A = kB c2B
or

(c A
2 0, c 0
B
2 ) c 2A k B
= = KD (16)
c 2B k A

7/10/07 3- 67
The ratio of the two Henrys law constants, k B/kA , is also a constant, which is known as the
distribution constant, or distribution coefficient, K D . Nernsts distribution law states that
for dilute solutions, a solute will divide itself between
two immiscible solvents to give a constant ratio of concentrations
(over a range of concentrations for which k A and kB are
constant).
The distribution law is the basis of solvent extraction
procedures. Distribution coefficients are also important, for
example, in the storage of gasoline, which is a mixture of many
hydrocarbons. The storage tanks must be vented to the air and
consequently water vapor can condense in the tanks. The water
is not significantly soluble and simply sinks to the bottom of the
tank, where it might be expected to cause no trouble. However,
the blending of gasolines, especially for winter driving, requires a
carefully controlled percentage of the lighter hydrocarbons, such
as butane and pentane. Because these hydrocarbons are more
soluble than the heavier hydrocarbons in water, the water layer
becomes richer in light hydrocarbons and the gasoline becomes
slightly depleted in the lighter, more volatile compounds. The
effect is to make the gasoline lack cold-weather starting
properties if the fractionation is not properly compensated for.

IDEAL SOLUTIONS. Henrys law states that, over the concentration range for which it is valid,
the fugacity, or vapor pressure, of a solute is proportional to the concentration of that solute.
The proportionality constant for each solute-solvent system must be determined by experiment.
In certain solutions the Henrys law constant takes on a particular value the value of the
fugacity, or vapor pressure, of the pure component when the concentration is expressed as a
mole fraction. Then, for the ith component,

(I. Soln.) f i = f io N i (17)

A solution for which each component obeys equation 17 is called an ideal solution. Equation 17
is called Raoults law. It will be shown later that Raoults law will necessarily apply for the
solvent (the major component) when a solution is very dilute, just as Henrys law must then apply
to the solute (the minor component).
One of the properties of an ideal solution is that there is no heat effect upon mixing the
components. Another property is that the volumes are additive.
(I.Soln.) V A+ B = V A + V B

Let vA and vB represent molar volumes of liquids A and B, and let V represent the volume of the
solution of A and B. Then, for an ideal solution comprising n A mol of A and nB mol of B,

7/10/07 3- 68
(I. Soln.) V = nA vA + nB vB (18)

PARTIAL MOLAL QUANTITIES. It is often convenient to have an


equation, for non-ideal solutions, similar to equation 18, by which
the total volume may be ascribed to the effective contributions
of the two components. These effective molar volumes (further
defined below) will be indicated
by V A and V B . Then

V = nAVA + nBVB (19)

for any solution of two components, A and B, whether the


solution is ideal or not. The equation may be extended to any
number of components.
If the number of moles of A is held constant and more B is
added to the solution, how much will the total volume increase?
Assume V A and V B are not significantly changed. 4 Then

dV = d (n A V A ) + d (n B V B )

but we have specified that nA is constant, so

(nA ) dV = VB dnB

and the partial molal volume is defined as


V
= VB (20)
n B nA

The partial molal volume (or partial molar volume) is the effective volume per mole. A graphical
interpretation of the partial molal volume is that it is the slope of the plot of total volume against
moles of the component added, at constant temperature and pressure, with the amounts of all
other components held fixed (Figure 3).
Even for an ideal solution, we cannot write an equation for free energy comparable to
equation 18.
G n1G1 + n2G2 + n3G3 + ...

because the mixing process introduces an increase of entropy that decreases the free energy. It is
particularly important, therefore, to define an effective, or a partial molal free energy. For the i th

4
It will be shown later that even this restriction is not important, because the changes
would be compensating.

7/10/07 3- 69
component of a solution,
G
Gi = (21)
ni T,P,n

This definition shows that the effective free energy per mole of the i th component is the rate of
change of the total free energy of the solution upon addition of a small additional amount of the i th
component, holding temperature, pressure, and the amounts of all other substances constant. As
in equation 19, one can write, for any solution,

G = n1G1 + n2 G2 + n3G3 + (22)

The partial molal free energy includes correction terms for the heat of mixing as well as the
entropy of mixing. One should therefore employ the partial molal free energy, or the effective
free energy per mole, in all equations. For a pure substance the partial molal free energy is
identical with the molal free energy.
The significance of the partial molal free energy in chemical thermodynamics cannot be
overemphasized; it is the necessary key to the treatment of all solutions, whether solid, liquid, or
gaseous. And, inasmuch as it includes pure materials as a special case, it provides a single
function for the thermodynamic treatment of all substances. A necessary condition for physical
equilibrium is that, for each substance, the partial molal free energy must be the same in every
phase. Furthermore, chemical reactions can occur only in a manner that will minimize the total
free energy (for the isothermal process), which is a sum of partial molal free energies, each
multiplied by the number of moles (equation 22). Thus Gi is a measure of the potential reactivity
of the ith species. Because of this particular importance, the partial molal free energy is often
called the chemical potential, and given the symbol i.

Chemical potential Gi = i

The free energy of a solution depends on the number of moles of each of the constituents, on
the temperature, and on the pressure. Therefore the change in free energy arising from changes in
any of these quantities can be written as the sum

G G G G G
dG = dn1 + dn2 + dn3 + + dP + dT
n1 n2 n3 P T

It is to be understood that each of the partial derivatives is evaluated with the other variables held
constant. Substitution of equation 21 and of equations 22 and 23 of Chapter 2 puts this equation
into the form

dG = G1dn1 + G2 dn2 + G3 dn3 + + VdP SdT (23)

This is a more general form of equation 21, Chapter 2.

7/10/07 3- 70
Equation 22 arises from the definition of partial molal free energy, and therefore is
completely general. It can be differentiated, term by term, to give

dG = G1dn1 + n1dG1 + G2 dn2 + n2 dG2 + G3 dn3 + n3 dG3 + ... (24)

Subtraction of equation 23 from equation 24 gives the Gibbs-Duhem equation

0 = n1dG1 + n2 dG2 + n3 dG3 + VdP + SdT (25)

which relates the changes in partial molal free energies to changes of the amount of any
constituent, changes of temperature, or changes of pressure all evaluated, of course, at some
particular concentration, temperature, and pressure. An important variation of this equation is
obtained by assuming constant temperature and pressure:
(T , P ) n1 dG1 + n2 dG2 + n3 dG3 + = 0 (25a)

This shows explicitly that the extra terms appearing in equation 24 (beyond equation 23) dont
add anything they add up to zero. But more important, equation 25a shows clearly that the
sum of changes in free energy terms vanishes; the changes compensate. A similar derivation for
partial molal volumes would justify the earlier assumption about compensating changes for
different components.

RAOULTS LAW. The addition of a small amount of solute to any solvent will, of course, increase
(from zero) the vapor pressure of the solute. It will also decrease the vapor pressure of the
solvent below the value for the pure solvent. The solute vapor pressure is given by Henrys law
(equation 15); the solvent vapor pressure is given by Raoults law, which says that the vapor
pressure of the solvent is proportional to the mole fraction of the solvent.
From equation 25 we can derive the relationship between the fugacity or vapor pressure of
the solvent and the composition of the solution. Substitution of equation 7a into equation 25a
gives, for three components,

n1 d ln f1 + n2 d ln f2 + n3 d ln f3 = 0

Dividing by n1 + n2 + n3 gives

(T,P) N1 d ln f1 + N2 d ln f2 + N3 d ln f3 = 0 (26)

For N2 and N3 very small, equation 15 requires that f2 /N2 = df2 /dN2 . Therefore,

(N2 0) N2 d ln f 2 = N2
df2 dN
= N2 2 = dN2 (27)
f2 N2

and similarly for the third component. Therefore,

7/10/07 3- 71
(N2 0, N3 0) N1d ln f1 + dN2 + dN3 = 0 (28)

and integration from pure solvent (N2 = N3 = 0) to dilute solution (N2 and N3 very small) gives 5

f1
N1 ln + N2 + N3 = 0
f1o
The logarithm can be expanded by means of the approximation

ln (1 - x) = - x

for x << 1.
f1 f1o f1 f1o f1
ln = ln
1 =
f1o f1o f1o

With this substitution, after multiplying through by f 1o,

N1 f1 N1 f1o + N 2 f1o + N 3 f1o = 0

Divide by N1 and rearrange, writing N2 for N2 /N1 and N3 for N3 /N1 because N1 is practically one in
the dilute solution. (This approximation, like others in the derivation, ignores terms of the order
of N22.)

f1 = f1o(1 - N2 - N3)

The sum of the mole fractions is, from the definition, 1. Therefore,
(N1 1) f1 = f1o N1 (29)

This is Raoults law, which must apply for the solvent in a very dilute solution. It gives the
fugacity of the solvent, in the solution, in terms of the fugacity of the pure solvent, f 1o, at the same
temperature and pressure. The notation (N 1 !61) at the left means that the equation is valid when
N1 is very nearly equal to 1, just as the notation (N2 !60) means N2 has been required to be very
nearly equal to zero.
Raoults law applies for nearly pure liquids, for which the vapor pressure is not necessarily
small and therefore the fugacity is not in general exactly equal to the vapor pressure. However,
the ratio of the fugacities will be equal to the ratio of vapor pressures when f 1 . f 1o and P 1 . P 1o.

5
N1 may be considered as constant during the integration, because it is always nearly 1.
Alternatively, equation 28 may be expressed in terms of n 1, n 2, and n 3 and the integrated equation
divided by N1 to give the same result.

7/10/07 3- 72
f1 P1
o
=
f1 P1 o

Thus Raoults law may be written in the more common form,

(N1 1) P1 = P1o N1 (29a)

An ideal solution is defined as one in which all components obey Raoults law at all
concentrations.
(I.Soln. ) fi = fio Ni (29b)

Because fi and fio may be quite different, fi /fio may be quite different from Pi /Pio . That is, the non-
ideality correction for the vapor may be quite different at different values of the pressure of the
gas (Pi vs. Pio). If the vapors can be assumed to be ideal, then

(I.Soln.,IdealVapors) Pi = Pi o N i (29c)

From equation 29b it follows that an ideal solution also has


additive volumes (equation 18) and additive enthalpies (no heat of
solution).

OSMOTIC PRESSURE. If a solution can exchange solvent with a


reservoir of pure solvent, at the same temperature and pressure,
solvent will necessarily pass from the pure solvent to the solution
because the fugacity of the solvent in the solution, f 1 = N 1f 1o, is
less than the fugacity of the pure solvent, f 1o.

f1 = N1 f1o < f1o

No finite amount of dilution of the solution by the solvent can


achieve equilibrium.
Equilibrium can, however, be reached if the fugacity, f 1, of
solvent in the solution is increased by an increase of pressure on
the solution (but not on the pure solvent reservoir). The
additional pressure required, which may be quite large, is called
the osmotic pressure.
One effective experimental arrangement for demonstrating
osmotic pressure is shown in Figure 4. The solvent, in the

7/10/07 3- 73
beaker, passes through the membrane (called semipermeable because it allows the solvent
molecules to pass, but not the solute molecules, which typically are larger). This dilutes the
solution in the thistle tube and forces the solution up the stem, increasing the hydrostatic pressure
on the solution.
Consider a system in which there is pure solvent, at atmospheric pressure, on both sides of a
membrane. Addition of solute to one side, to form the solution, lowers the fugacity of the solvent
by dilution, but an increase of pressure on the resultant solution can increase the fugacity to its
original value. The free-energy changes caused by addition of the solute and by the change of
pressure will be
G1 G1
dG = dn + dP
n 2 P n2
1 2
P

and the sum of these terms must be zero, to maintain equilibrium with the pure solvent phase.
The free-energy values should be understood to be the effective, or partial molal, free energies in
these equations. Substitution of equation 7a into the first term and equation 22 of Chapter 2 into
the second, representing the volume of the solvent by its partial molal value, V 1 , gives

RTd ln f1
dN2 + V1dP = 0 (30)
dN2

The first term is reduced with the aid of equation 28.

( N 2 0) d ln f1 dN
dN 2 = 2 = dN 2 (31)
dN 2 N1

and therefore
RTdN 2 + V1 dP = 0
or

RT
dP = dN2
V1
Integration from N2 = 0 to N2 = N2 (a very small value) gives the expression for the osmotic
pressure,

(N2 0) P =
RT
N2 (32)
V1

The pressure produced can be quite large, even for dilute solutions. The molar volume of
water at 25oC is 18 x 10-6 m3 and RT is 8.314 x 298 = 2.48 kJ/molK. Thus if N2, the mole

7/10/07 3- 74
fraction of impurity in the water, is only 10 -3, the osmotic pressure is 1.38 x 105 Pa = 1.36 atm or
1032 torr. The effectiveness of osmotic processes is often limited by the poor selectivity of
available membranes, usually based on size differences between solute and solvent molecules.
Equation 32 can be easily remembered by its similarity to the ideal-gas law.
(P)V1 = N2 RT looks like PV = nRT. Note, however, that the volume, V 1 , refers to the
solvent, whereas N2 is the number of solute molecules in that volume. 6
Osmosis may be considered to be simply a diffusion of solvent molecules from a more
concentrated (in solvent) solution to a less concentrated (in solvent) solution. It is of great
importance to life processes. The direction of diffusion of water can be reversed by exposure of
plant or animal cells to solutions more concentrated in solute than are the cell fluids (e.g., by
over-fertilization). Reverse osmosis has been applied to the problem of producing fresh water
from salt water. When the salt water is compressed beyond the osmotic pressure, water flows
through a membrane, leaving the salt behind.

CHANGE OF BOILING POINT OR FREEZING POINT. Addition of sugar to water will raise the boiling
point but lower the freezing point of the water. Addition of gasoline to water will lower the
boiling point but leave the freezing point essentially unchanged. Addition of gold to silver will
raise the freezing point of the silver, although addition of silver to gold lowers the freezing point
of the gold. These effects, which may seem to have no consistent pattern, are all predictable by a
single equation, knowing only certain properties of the pure substances and the mutual
solubilities.
An impurity added to a substance at a phase transition point will usually go preferentially into
one phase or the other. This lowers the fugacity of the major component in the phase that has
become impure and upsets the equilibrium.
Solvent is present in both phases and it is a necessary condition for the equilibrium that the
free energy of the solvent (as well as solute) should be the same in the two phases, both before
and after addition of solute. After addition of solute to one phase, if equilibrium is re-established
by changing the temperature of the two phases, the condition for final equilibrium is that the sum
of concentration and temperature effects on the free energy of the solvent in the impure phase
should equal the temperature effect on the free energy of the solvent in the pure phase.

dG1 = dG1o

Substituting fugacities
(equation 7a), ln f 1 ln f 1 ln f 1o
dN 2 + dT = dT
N 2 T T N 2 T N2

6
A pseudo-derivation is as follows. If P1o V1 = N 1o RT for pure solvent, P1V1 = N 1 RT for
the solution, and V1 can be assumed equal for both, then
(P1o P1 )V1 = (N1o N1 )RT or PV1 = N2 RT . (But of course the ideal-gas law is not really
applicable to solutions.)

7/10/07 3- 75
For example, if sugar is added to water at the boiling point, the change in fugacity of the liquid
water is given by the left-hand side and the change in fugacity of the pure (that is, sugar-free)
water vapor is given by the right-hand side. From equation 31 the first term is just - dN 2, and
from the exact Clausius-Clapeyron equation (12),

d ln f 1 H * H d ln f 1o H 1 * H 1o
= 1 2 1 and =
dT RT dT RT 2

in which H1* is the enthalpy of the solvent vapor at very low pressure, H1 is the (partial molal)
enthalpy of the solvent in the impure phase at the pressure of the equilibrium, and H 1o is the
(partial molal) enthalpy of the solvent in the pure phase in equilibrium with the impure phase.
Substitution of these expressions gives

H 1 * H 1 H 1 * H 1o
dN 2 + =
RT 2 RT 2

and, combining terms,

dN2 =
(
H1 * H1 H1 * H1o
dT
)
RT 2
Ho H
= 1 2 1 dT
RT
This relates the change in temperature, from the normal equilibrium temperature for the pure
substance, to the change in mole fraction of impurity in the substance (in one phase only).
Because N2 = N2 - 0 = N2, the equation may be rewritten
H 1o H 1
(N 2 0) N2 = T (33)
RT 2

It is clear that T, which is the change in boiling point or freezing point, must have the same sign
as H1o - H1, which is the enthalpy for the solvent going from the impure phase to the pure phase.7
For example, if a solution of sugar and water is boiled, the pure phase is the water vapor and
thus H1o is the enthalpy of water vapor at the prevailing pressure; the impure phase is the liquid
solution and so H1 is the (partial molal) enthalpy of the water in the sugar solution; and H1 o - H1 is
therefore the heat of vaporization of water, which is positive. It follows that the boiling point of a
sugar-in-water solution must be higher than the boiling point of pure water, at the same pressure.

7
As a mnemonic device one may remember that scientists, and indeed all moral
individuals, strive for greater purity. Thus H is the thermal energy absorbed in going to the
purer phase.

7/10/07 3- 76
If the solution of sugar and water is cooled to the freezing pint, pure ice will separate out. Then
the pure phase is the ice and so H1o is the enthalpy of the ice, H1 is the partial molal enthalpy of
the water in solution, H1o - H1 is the heat of freezing (- 334 J/g), and T will be negative.
If CCl4 is mixed with water it will not dissolve, but the vapors mix. Then N2 will be the mole
fraction of CCl4 in the water vapor and H1 o - H1 will be the enthalpy of the (pure) liquid water
minus the enthalpy of water vapor; this H is negative and the boiling point of the water is
accordingly decreased. The boiling point of CCl 4, or of any other liquid immiscible with water,
will be similarly depressed by the addition of water, a phenomenon known as steam distillation.
An alternative expression is

H 1 T
N 1 = (34)
RT 2

where N1 is the difference in mole fraction of the major component (N1A - N1B ) and H1 is the
difference in enthalpy of the major component (H 1A - H 1B).

THE IMPORTANCE OF COLLIGATIVE PROPERTIES. The equations for lowering of vapor pressure
of the solvent (Raoults law), for osmotic pressure, and for change of boiling point or freezing
point, share a common characteristic. Each of these expressions involves the mole fraction of the
solute or, equivalently, the mole fraction of the solvent, and is totally independent of any of the
individual properties of the solute. Within the range of dilute-solution laws, the effects calculated
for any given solvent depend solely on the properties of that solvent and on the total number of
molecules of solute added, even if there is a mixture of several kinds of solute molecules. That is
why the quantities measured (vapor pressure lowering, osmotic pressure, and change of boiling or
freezing point) are called colligative properties.
Measurement of a freezing point lowering, for example, will permit calculation of the number
of moles of solute, so that, knowing the number of grams of solute added, an average molar mass
can be found. This is an effective means of showing whether a solute dissociates, and into how
many parts. The colligative properties are also very important for determining molar masses in
polymeric materials.
The limitation to very dilute solutions is not a severe handicap in practice. Measurements
can be made at several concentrations and the values extrapolated to zero concentration. The
dilute-solution laws provide assurance that the extrapolation will be a straight line at the low
concentration end. Also, in many applications the important question is whether the solute is or is
not dimerized, or dissociated. An approximate value, at moderate concentrations, can provide the
answer.

Phase Diagrams
The dilute-solution equations provide important information concerning behaviors of both
solvents and solutes when solutions approach infinite dilution, but they lack information that may
be required. They do not predict actual melting points or boiling points of pure materials, nor do
they predict values of the Henrys law constants. Only very general characteristics are predicted

7/10/07 3- 77
for concentrated solutions. Such information is usually best presented in graphical form. Only a
brief introduction is possible here, but it will provide a guide to the meaning of phase diagrams
and how to interpret them.
By a phase we will mean any generally homogeneous region, whether vapor, liquid, or solid.
Because vapors always mix, there can be only one vapor phase. Some liquids are miscible, others
only partially so, and some are essentially immiscible. Solids generally do not mix significantly,
although some solid solutions are known.

GIBBS PHASE RULE. You probably recall the general axiom from mathematics for solution of
equations in more than one unknown. If there are n variables, or unknowns, there can be a
unique solution if and only if there are n conditions, or equations. Two unknowns require two
equations, and so forth. More than two (independent) equations give no possible solutions.
Fewer conditions give an infinite number of possible solutions. The number of conditions
permitted but unspecified is called the number of degrees of freedom.
For example, if a salt may have three distinct hydrate forms (each with a different number of
water molecules incorporated into the crystal structure), then only a certain number of these may
co-exist in equilibrium. Fewer hydrates present imply one or more degrees of freedom, whereas
more hydrates than the allowed number will cause some of the forms to disappear as water
molecules are exchanged between them.
Let v represent the number of variables and e the number of equations relating them. Then
the number of degrees of freedom, f, is

f=v-e (35)

Each independent component8, c, contributes a concentration value in each of the p phases,


except that it is sufficient to know c - 1 concentrations in each phase. Thus

v = p(c - 1) + external variables

Typically pressure and temperature are external variables, so

v = p(c - 1) + 2

Each component must have the same free energy, or fugacity, in each pair of phases, giving
c(p - 1) equations, or conditions. Therefore the number of degrees of freedom (remaining) is

8
Usually the number of components is simply the number of things you would mix
together. Occasionally there is possible uncertainty. You might, for example, mix ammonia gas
and hydrogenchloride to form NH4Cl. Then, because there may be excess NH3 or HCl (but not
both), there would be two components. However, if you began with NH4Cl, the ratio of HCl to
NH3 is predetermined and there is only one component.

7/10/07 3- 78
f = p(c - 1) + 2 - c(p - 1)
=c-p+2 (36)

Figure 5. Each area


corresponds to a single
phase. The lines are the
common borders, and
thus represent two phases
in equilibrium (or three
phases at a point)..

which is Gibbs phase rule, for two external degrees of freedom.


You will recall from experience with drawing graphs that a point (or any finite number of
points) corresponds to no degrees of freedom. A line (or a finite number of lines) represents one
degree of freedom; setting one variable determines the other. Two degrees of freedom yield an
area (or more than one area).

ONE-COMPONENT PHASE DIAGRAMS. Water is a single component (c = 1), so the degrees of


freedom may vary from zero to two.

(c = 1) f=c-p+2=3-p (37)

A single phase (vapor, liquid, or solid) is therefore represented by an area. Two phases (solid-
liquid, solid-vapor, or vapor-liquid) are restricted to a line. Three phases yield a single point,
called the triple point (solid + liquid + vapor). The phase diagram (with T on the vertical axis, to
match following diagrams) is shown in Figure 5.
You can calculate the slopes of the three lines from the Clapeyron and Clausius-Clapeyron
equations, knowing the heats of fusion and of vaporization or sublimation and the densities of the
phases. The upper end of the liquid-vapor line simply stops above the critical point the
distinction between liquid and vapor disappears.

TWO-COMPONENT PHASE DIAGRAMS AT CONSTANT PRESSURE. We could not represent a single


phase with two components on a two-dimensional page if we kept both temperature and pressure
as external variables. We consider here the restricted example where pressure is maintained
constant (e.g., at atmospheric pressure). Then Gibbs phase rule tells us the degrees of freedom
are, again, with c = 2,

(c = 2; const P) f=c-p+1=3-p (38)

It follows that wherever there is a single phase present, that phase must be represented by an area
(two degrees of freedom). If there are two phases present, the compositions are restricted to a
line (or sometimes more than one line, but still one degree of freedom). Three phases can coexist

7/10/07 3- 79
only at a point, or a set of
points, with no remaining
degrees of freedom.
Accordingly, an area can
represent only a region in
which there is a single phase,
of variable composition.
There are two apparent
exceptions. First, the
solubility of one solid in
another is typically so

Figure 3.6. Benzene-toluene equilibrium (liquid-vapor, at constant pressure). a) Calculated equilibrium


states. b) Phase diagram. The area between the curved lines is not part of the equilibrium phase diagram.
There can be no phase of such temperature and composition.

small that the area of a pure solid will look like a single, vertical line. Second, when two
phases, solid or liquid, are in equilibrium, those two phases will have different compositions and
therefore different areas, bordered by lines that are separated from each other, as in the benzene-
toluene diagram of Figure 6. Each line is simply the border of its respective area. If the overall
composition falls between those lines, or areas, the material will split into variable amounts of the
left and right areas (phases). The region between phases has been given the unfortunate label of a
two phase region; a better description is a forbidden area. There can be no phase present with
a composition falling in such a forbidden area. The relative amounts of the left phase and right
phase are given by the lever-arm rule. If C L and C R are the compositions (as percentages of either
phase say the left-hand phase) and C is the overall composition of the mix at some point, the
relative amounts (masses) of the left-hand and right-hand phases are given by

m1 (C - CL ) = m2 (CR - C) (39)

Similar behavior is shown by two immiscible solids or two immiscible liquids. Carbon
tetrachloride and water liquids split into two layers, with very little solubility of either liquid in the
other. At higher temperatures, however, the vapors mix freely. Thus the basic phase diagram is
as shown in Figure 7a. There are three possible phases: the vapor, at higher temperatures, liquid
CCl4 at room temperature and below, and liquid water. The two liquid phases can co-exist, but
there is no liquid phase with composition falling in the general range of 95% CCl 4/5% water to
95% water/5% CCl4 (where the numbers are illustrative only, and depend somewhat on
temperature). Figure 7a is somewhat difficult to interpret, so it is customary to add a horizontal
line as a marker, to dramatize the lowest vapor temperature of the mixture, as in Figure 7b.
(There can be no equilibrium across a horizontal line, which would connect points at different
temperatures.) Also, if the diagram represents solid-liquid equilibrium, as for silver and copper in
Figure 7c, it was recognized long before Gibbs work of the late 19 th century that when the liquid

7/10/07 3- 80
solution is cooled, either silver or copper will precipitate out first (depending on whether the
composition lies to the left or right of the low point, called the eutectic). The composition of the
remaining liquid accordingly moves toward the eutectic composition. In the last stage of
precipitation, or freezing, the mixture that comes down is a fine mix of two phases, silver and
copper, with the overall composition of the eutectic. Accordingly, a vertical line was drawn
below the eutectic point to represent this eutectic mixture two distinct phases, but with a
distinctive appearance. Over decades, the vertical line was changed to a dotted line (to

a b c
Figure 3.7. a) Bare-bones phase diagram for two immiscible phases with miscible high-
temperature phase (e.g., CCl4 and H2O, liquid-vapor, or Silver-Copper, solid-liquid). b) The
horizontal line guides the eye. c) Early diagrams showed the eutectic mixture as if it were a
separate phase, or (later) as a dotted line.

indicate there is no phase with composition falling in this broad forbidden region) and more
recently the vertical line is largely omitted.
Examining a large number of such phase diagrams, the two patterns shown here will be
recognized as appearing in combination. Often a solid compound forms, so the diagram can
conceptually be divided in two, element one plus compound on the left and compound plus
element two on the right. There may be regions of solubility, resembling the benzene-toluene
diagram, within a larger phase diagram. Some compounds may spontaneously decompose as the
temperature rises, yielding separate phases. The variety is seemingly endless. But almost all can
be recognized as relatively simple combinations of these two basic patterns.

Problems
1. For the transformation of graphite to diamond at 25 oC, G = 2.9 kJ/mole. The densities are
2.25 and 3.51 g/cm3. What is the minimum pressure required to make diamond
thermodynamically stable at this temperature?
2. At 1114oC the vapor pressure of Ni was observed to be 7.50 x 10-8 torr and at 1142o C it was
14.33 x 10-8 torr. What is the heat of vaporization of nickel in this range?
3. Is the fugacity of ice increased or decreased by exerting pressure on the ice? Which will show
the greatest change in fugacity for a pressure increase of 10 atm: ice, or liquid water? Explain.
4. The vapor pressure of a liquid compound, Y, obeys the equation ln P = a + bT -1, with P the
pressure in atm and T in kelvin.

7/10/07 3- 81
a. What is the boiling point of Y?
b. What is the heat of vaporization at the normal boiling point?
5. Find the mole fractions of the several components if 24 g of methane, CH 4, is mixed with 7 g
of CO, 14 g of N2, and 8 g of He.
6. A saturated solution of benzoic acid, C6H5COOH, contains 0.32 g in 100 g of water. The
density of the solution is 1.3 g/cm3. What is the molarity of the solution?
7. Find the mole fraction, molarity, and molality of iodine when 0.10 g of iodine is dissolved in
100 g of CCl4 (density 1.59 g/cm3).
8. a. What is the molarity of water in pure water?
b. What is the molality of water in pure water?
c. What is the molarity of CCl4 in pure CCl4 (density 1.59 g/cm 3)?
d. What is the molality of CCl4 in pure CCl4?
e. What is the molarity of an ideal gas at 1 atm and 25 oC?
9. A solution contains, by weight, 40% water, 35% ethanol (C 2H5OH), and 25% acetone
(CH3COCH3).
a. Find the mole fraction of each component.
b. Find the molality of the ethanol and acetone if water is considered to be the solvent.
c. Find the molality of the water and acetone if the ethanol is considered to be the solvent.
10. The solubility of CCl4 in water at room temperature is about 0.9 g/L and the vapor pressure
of CCl4 is about 100 torr. What would be the equilibrium vapor pressure of CCl4 above a beaker
containing CCl4 covered with a layer of water?
11. Water in equilibrium with air (20.9% O2, 79.1% N2 by pressure, volume, or mole fraction) at
0oC contains 1.28 x 10-3 mole of air per liter, of which 34.91% is O 2. Calculate the Henrys law
constant
a. for O2
b. for N2
12. Is the Henrys law constant for benzene in water large or small?
13. The distribution coefficient for mercuric bromide, HgBr 2, between water and benzene is 0.90.
That is, with concentrations expressed as molarities, C(HgBr2 /H 2 O) = 0.90.
C(HgBr2 /C 6 H 6 )
An aqueous solution contains 0.010 M HgBr2. 50 ml of this solution is to be extracted with 150
ml of benzene.
a. What fraction of the HgBr2 is left in the aqueous phase if the extraction is made with one
150 ml portion of benzene?
b. What fraction is left if the extraction is made with three successive 50 ml portions of
benzene?
14. Over a certain range of concentrations, the volume of a solution containing m moles of NaCl
in 1 kg water has been found to be, in ml,
V = 1002.935 + 16.670 m + 1.636 m3/2 + 0.170 m2
Find the partial molal volume of
a. NaCl in a 2 m solution
b. water in a 2 m NaCl solution
c. pure NaCl (density 2.165 g/cm3)

7/10/07 3- 82
15. The vapor pressures of benzene and toluene, at 27 oC, are 120 and 40 torr. Assuming an ideal
solution, what is the composition of the vapor above a solution containing 100 g of benzene
(C6H6) and 150 g of toluene (C6H5CH3 )?
16. In your work for Adulterated Chemicals, Inc., you have isolated a new antibiotic through a
lengthy series of extractions, biological tests, and so forth. A few milligrams are available, and by
the ultra-centrifuge method it has been found that the molar mass is 10,000. It is desired to check
this by another method. For a 1% by weight solution of the substance in water, calculate
a. the freezing-point depression
b. the boiling-point elevation
c. the change in vapor pressure at 25oC
d. the osmotic pressure, at 25oC, in cm H2O. (The density of mercury is 13.6 g/cm3 .)
Which method would you recommend?
17. Boiling occurs in a solution when the sum of the vapor pressures of the components is equal
to the atmospheric pressure (or the applied pressure), but these vapor pressures are lower than for
the pure materials (cf. Raoults law, equation 29c). If two liquids are immiscible, each exhibits its
own vapor pressure (in each phase), and steam distillation occurs when the sum of the vapor
pressures is equal to the atmospheric pressure.
a. Calculate the vapor pressure of pure water at 99 oC.
b. What vapor pressure must a compound, immiscible with water, have to steam distill at
o
99 C when the atmospheric pressure is 745 torr?
c. What would be the composition of the vapor in such a steam distillation?
18. A solution of 3.795 g of sulfur in 100 g of CS2 boils at 46.66oC. The boiling point of pure
CS2 is 46.30oC and the heat of vaporization of CS2 is 26.8 kJ/mol. From this experimental result,
what is the probable formula of the sulfur molecule in the solution?
19. Show that if two phases (A and B) in equilibrium are both slightly impure, the resultant
change in temperature (change in melting point or boiling point, relative to pure phases), T, is
given by
H 1T
N 1 =
RT 2

where N1 is the difference in mole fraction of the major component (N1 = N1 A - N1 B ) and H1 is
the enthalpy of transition ( H 1 = H 1A H 1B ) .

7/10/07 3- 83
4 Chemical Equilibrium
Chemical equilibrium is one of the most important applications of thermodynamics. The
theory follows very simply from the equations already derived. Setting up the necessary
equations in practice requires only an understanding of the meaning of a chemical equation. Most
mistakes arise in the elementary process of counting the molecules or moles of reactants and
products.
For some systems the solution of the mathematical equations requires some skill and,
especially, some understanding of the physical implications so that reasonable assumptions and
approximations can be made. For the types of problems considered here, the primary requirement
is to follow a simple, methodical procedure in finding the appropriate equations, and then to insert
numerical values into these equations. Solution of the equations is then quite easy.

Free-Energy Changes in Chemical Reactions


A completely general chemical reaction may be represented in the form

aA + bB -----6 cC + dD

This should be interpreted as follows: Under certain conditions of overall temperature, pressure,
and concentrations of A, B, C, and D (not yet specified here), a moles of substance A react with b
moles of substance B to form c moles of substance C and d moles of substance D. Although
some of the reactants (A and B and perhaps others) are consumed and some products (C and D
and perhaps others) are formed, it is assumed that there is no change in temperature, in pressure,
or in concentrations of any of the substances as a consequence of the chemical reaction.
The free-energy change, if the reaction proceeds as written, will be

G = n G
products
i i n G
reactants
j j

= cGC + dGD aG A bGB (1)

The values of the partial molal, or effective, free energies on the right-hand side depend on the
states of these compounds, including the temperature, the pressure, and the individual
concentrations. Thus G may have any value, depending on the states of reactants and products.
Remember that, for purposes of the calculation, the states of reactants and products are
unchanged by the reaction. The temperature, pressure, and concentration of each substance is the
same before and after the reaction. The reaction may involve a negligible fraction of the
substances present, or reactants may be added and products removed as the reaction proceeds.
Tables give G = Go for the reaction occurring with all reactants and products in their
standard states.
Go = cGCo + dGDo aGAo bGBo (2)

7/10/07 4- 84
Subtraction of equation 2 from equation 1 gives
( ) ( ) ( ) (
G Go = c GC GCo + d GD GDo a GA GAo b GB GBo ) (3)

At this point it is convenient to substitute the fugacities (equation 7a, Chapter 3) for the free-
energy terms.

G i = RT ln f i + B i (T )
and G i o = RT ln f i o + B i (T )
Therefore,
fi
Gi Gio = RTln (4)
fi o

Substitution of equation 4 into equation 3, for each of the substances, gives

f f f f
G Go = RT c ln Co + d ln Do a ln Ao b ln Bo
fC fD fA fB
The ratio of the fugacity of any substance, f i, to the fugacity of the same substance in its standard
state, fio, is called the activity, ai .

fi
ai = (5)
f io

It follows from the definition that activity is always dimensionless and without units. Substitution
of activity for the ratio of fugacities and replacement of n ln x by ln x n gives

(
G G o = RT ln a Cc + ln a Dd ln a Aa ln a Bb )
which can be further simplified to the form

aCc a Dd
G = G + RT ln a b
o
(6)
a A aB
It will be convenient to write this in the form

G = G o + RT ln Q (a ) (7)

defining Q (a) to be
aCc aDd
Q(a) = (8)
a Aa aBb

7/10/07 4- 85
or the corresponding form appropriate to the particular reaction.
Equation 6 is a very important result with obvious physical interpretation. If all substances
(reactants and products) are in their standard states, all activities are, by definition, equal to 1.
Then Q (a) = 1, ln Q (a) = 0, and G = Go. The correction term depends on the value of Q (a),
which depends on the individual activities and therefore on the fugacities, or on the pressure and
concentration of each substance at the given temperature.

Standard States
The choice of standard states affects the value of G o as it appears in tables, and determines
the relationship between the state of a compound and its activity. There are certain conventions
that are customarily followed, so that it is generally considered unnecessary to explain in detail
the choice of standard states that has been made. Unless the conventions are understood,
therefore, it may not be possible to apply thermodynamic values found in the literature. Standard
states are chosen to give a convenient form for the activity.

GASES. For an ideal gas, fugacity is equal to pressure. We choose the standard state to be 1 atm
pressure. Then f = P, fo = Po = 1 atm:
f P(atm)
a= o
= = P(atm) (9)
f 1 atm

That is, the activity, which is a dimensionless ratio, is equal to the numerical value of the pressure,
with the pressure expressed in atmospheres. This simple relationship can be extended to cover
real gases by the introduction of an activity coefficient, (Greek gamma), that will equal 1 if the
gas is ideal, and will differ from 1 as the gas deviates from ideality.

a = P (10)

The activity coefficient defined in this way has units of atm -1. Equation 10 is applicable to any
gas, real or ideal.1 If the gas is non-ideal, it is necessary to know the value of .

LIQUIDS AND SOLUTIONS. The major component of an ideal solution has a fugacity given by
Raoults law

1
More generally, the standard state of any gas is taken as the ideal gas at 1 atm pressure;
that is, it is a fictitious state with properties obtained from the properties of the actual gas, at
sufficiently low pressures that the gas is ideal, extrapolated to 1 atm pressure using the equations
for behavior of an ideal gas. The numerical value of the activity coefficient is the ratio of the
fugacity of the real gas to that of the idealized gas. The dimensions are reciprocal pressure.

7/10/07 4- 86
f1 = f1o N1
in which f1o is the fugacity of the pure solvent. We choose the pure solvent as the standard
state so that f1o is the fugacity in the standard state. Then
f1
a1 = = N1 (11)
f1o

Thus the activity of the solvent is equal to its mole fraction, in the ideal solution, and the activity
of any pure liquid is 1. Not all solutions are ideal, so it is convenient to allow for deviations from
Raoults law by introducing an activity coefficient, as for gases, defined by the equation

a1 = 1 N 1 (12)
This activity coefficient is dimensionless. It will be 1 for a pure liquid, for a solvent in an infinitely
dilute solution, or for a component of an ideal solution. It may differ appreciably from 1 when the
fugacity of the liquid in solution deviates from Raoults law.
If a solute obeys Henrys law, the fugacity is proportional to concentration and the standard
state of the solute may be chosen such that f 2o = k, the Henrys law constant (even though such a
standard state may be physically unattainable). Then
f2 f2
a2 = = = c2 (13)
f 2o k

with |c2| representing the numerical value of c2 . For any particular concentration, the value of the
activity depends on the units chosen for expressing concentration. Common choices are molarity,
molality, or sometimes mole fraction (especially when the choice between solute and solvent is
ambiguous). Activity is dimensionless but the activity coefficient has dimensions of inverse
concentration. For example, if c2 is in mole per liter, 2 has units of liter per mole. In any case,
a2 = 2 c2

Molality has some advantages in accommodating changing temperature. Nevertheless, in


subsequent discussions we will assume the concentration of solute is expressed as molarity. For
dilute water solutions the difference between molarity and molality is negligible.
Concentrated solutions may show large deviations from Henrys law, as a consequence of
large intermolecular forces. This may also occur with uncharged solutes, leading, for example, to
a maximum-boiling or minimum-boiling solution, known as an azeotrope. It is also expected for
multiply charged ions (which typically form complex ions of smaller charge). In dilute solutions,
activity coefficients of ions are less than 1, approaching 1 at infinite dilution. Activity coefficients
of ions in concentrated solutions (on the order of 1 M) may increase and become greater than 1.
Strong electrolytes fully dissociate in aqueous solution, so the concentration of undissociated
electrolyte may be assumed to be vanishingly small. We can eliminate consideration of the
undissociated electrolyte by defining its activity to be equal to the product of the activities of the
ions into which it dissociates. Then the equilibrium constant for the dissociation is 1, G o for the

7/10/07 4- 87
dissociation is zero, and the irrelevant quantities disappear from the calculations.

SOLIDS. The fugacity of a solid is very nearly constant, at a given temperature, independent of
the applied pressure (except for extreme pressures). We choose this normal value of the fugacity
as f o, so the solid has an activity of 1. Appreciable deviations from unit activity for a solid may
be expected only under very high pressures, when impurities are present (as in mercury-metal
amalgams), when the solid is severely strained, or when the solid exists as extremely small
particles for which surface effects cannot be neglected.

VARIATIONS AND LIMITATION. Other choices of standard states are possible and may be
convenient for special circumstances. For example, the standard state of a gas may be taken as a
concentration of 1 mol/L rather than at 1 atm. Such a choice seldom, if ever, appears in tables of
standard free energies, but may occasionally be found, especially in introductory discussions, for
applications to chemical equilibria.
There is an important distinction between the choices of standard states for enthalpy and for
free energy. Both are commonly taken for 1 atm pressure and for the most stable phase under the
standard conditions. However, enthalpies are given for a fixed temperature, usually 25 oC (but
sometimes 18oC). Corrections for other temperatures are made by calculations involving the heat
capacity.

H (T2 ) = H (T1 ) + C P dT
T2

T1

On the other hand, the temperature dependence of the free energy is given by equation 23,
Chapter 2,
G
= S
T P

We have defined entropy only through S, and therefore subject to an arbitrary additive constant.
The change in free energy with temperature is thus also subject to an arbitrary additive constant,
and therefore is unknown.2 It is therefore necessary to choose the standard state for the
specification of free energy to be at the temperature of interest.

CALCULATIONS. We are now prepared to calculate free-energy changes for specific chemical
reactions for particular choices of conditions. Consider first the reaction for the synthesis of
ammonia at 25oC when each of the three gases is present at a partial pressure of 5 atm. The
standard free-energy change, Go, is - 16.64 J/mol (NH 3). Assume the gases are ideal.

2
It is possible to assign an arbitrary absolute value for entropy, but such methods always
omit some parts of S, and thus should be regarded as indeterminate. Note, however, that this
limitation does not apply to finding the change in G (as, for example, in finding the change in
Greaction) with change in temperature. This change depends on S, which can be measured. The
temperature dependence of fugacity can also be found.

7/10/07 4- 88
N2 + 3/2 H2 ------6 NH3

a NH 3
G = G o + RT ln
(a ) (a )
N2
1/ 2
H2
3/ 2

PNH 3 5
= G o + RT ln = 16 ,640 + 2, 478 ln
(P ) (P )
N2
1/ 2
H2
3/ 2
25
= 20 .62 kJ/mol (NH 3 )

The standard free-energy change for the decomposition of Ag 2O at 25 oC is 10.84 kJ/mol.


The free-energy change for Ag2O in air (20% O2) will be slightly different:

Ag2O ------6 2 Ag + O2

(a ) (a ) 2 1/ 2

G = G + RT ln
o Ag O2

aAg2O

Activities of the solids are 1 and the activity of the oxygen is 0.20 (equal to the pressure in
atmospheres).

G = 10,840+8.314 x 298 ln (0.20) = 10,840 - 1,239 ln 5


= 8.846 kJ/mol

The free-energy change for the dissolution of CaF 2 in water at 25 oC, containing 0.001 M NaF
and 0.0001 M CaCl2 may be calculated as follows. The standard free energies of solid CaF2 and
of CaF2 as a solute are -1,162 and -1,106 kJ/mol. Thus Go for the dissolution process is 55.90
kJ/mol.

CaF2 -------6 Ca++ + 2 F-

Setting aCa++ = |cCa++| = 0.0001 and aF- = |cF-| = 0.001 gives

(
G = 55900 + 8.314 298 ln 10 4 10 3 )( ) 2
= 55900 + 2478 ln10 10
= - 1.150 kJ/mol
The dissolution of HCl gas, at a pressure of 0.10 atm, is somewhat less spontaneous if the
water contains 0.50 M HCl. The standard free energies of HCl(g) and HCl(aq) are - 95.27 and
- 131.2 kJ/mol, so Go = - 131.2 - (- 95.27) = - 35.9 kJ/mol, and (by chosen convention) a HCl(aq)
= aH+ aCl-.

7/10/07 4- 89
a H + aCl
G = G o + RT ln
a HCl ( g )

= 35.9 + 8.314 298 ln


(0.50)(0.50)
0.10
= 33.6 J/mol

Equilibrium Constants
At equilibrium, under constant temperature and pressure, the free-energy change is zero and
equation 6 becomes
ac ad
0 = G o + RT ln Ca Db = G o + RT ln Q (a )eq

a AaB eq
or
ac ad
G o = RT ln Ca Db (15)
a AaB eq

For any given reaction, the left-hand side, G o, is a constant, obtainable from handbook tables. It
depends only on (1) the choice of standard states, and (2) the temperature. Therefore, at any
given temperature the right-hand side of equation 15 and therefore Q (a) eq must also be a
constant, which is called the equilibrium constant, K (or K eq).

a Cc a Dd
Q (a )eq = a b = K
(16)
a AaB eq

This equation is really quite a remarkable result, and one that students are often unwilling to
accept in its full implications. It says that, despite changes of pressure, concentrations of
reactants or products, or the presence of other substances, the particular function of activities
represented by Q (a) can assume only one value at equilibrium, for a given temperature and
choice of standard states by which the activities are defined. 3 The value of Q (a) at equilibrium is
determined by equation 15, which can be rewritten as

3
Undoubtedly part of the difficulty (whether cause or consequence of the confusion) is
that many textbooks state that the equilibrium constant depends on pressure or on the presence of
other substances, but this is because they are not talking about the equilibrium constant but rather
the value of Q (c) at equilibrium, which is not necessarily a constant. The changes in the
equilibrium constant referred to are really changes in the activity coefficients, or changes in the
deviations from ideality. The true equilibrium constant (or thermodynamic equilibrium
constant) is rigorously constant!

7/10/07 4- 90
Go = - RT ln K (17)

This equation was derived without benefit of arbitrary assumptions, so it must be satisfied for all
chemical equilibria.
There is a strong tendency to misinterpret equation 17. The left-hand side refers to standard
state conditions, the right-hand side to equilibrium conditions, but these are not the same. It may
be helpful to think of the equation in an expanded form:

(G)std st - (G)equil = RT ln Q std st - RT ln Q equil

Now we recognize that (G)std st = Go, (G)equil = 0, Q std st = 1 and ln 1 = 0, and Q equil = K.
Therefore the equation may be written as

Go - 0 = 0 - RT ln K

which may be interpreted as saying that G o is a measure of how far the standard-state
concentrations (or activities) differ from an equilibrium point, and the equilibrium constant is a
measure of how much the equilibrium concentrations (or activities) differ from their standard-
state values (activities equal to 1). The minus sign shows that the left-hand (standard state) side
deviation from equilibrium is equal and opposite to the right-hand (equilibrium) side deviation
from standard states.
Application of the equilibrium constant is straightforward if each step is taken in turn.
Consider, for example, the problem of finding the equilibrium pressures of N 2 and H 2 formed by
decomposition of ammonia, initially at 2 atm, at 25 oC.

NH3 ------6 N2 + 3/2 H2

(a ) (a ) 1/ 2 3/2

K =
N2 H2

a NH 3
G o = 16 , 640 J/mol = RT ln K
K = 1 . 2 x 10 - 3
In a gaseous system, activities are equal to pressures in atm (assuming ideal gas behavior).
The pressures, in turn, are proportional to the numbers of moles so, for a given volume and
temperature, it is convenient to think of the chemical reaction in terms of pressures (rather than
just moles). Thus, if the initial pressure of NH 3 is 2 atm and x atm decomposes, the chemical
equation states that x/2 atm of N2 and 3x/2 atm of H2 will be produced, with 2 - x atm of NH3
remaining. The equilibrium condition requires that

(x / 2)1 / 2 (3x / 2)3 / 2 = 1.2 x 10 -3


(2 x )

7/10/07 4- 91
This is a quadratic equation, with the solution x = 0.0428 atm. The equilibrium condition may
now be more explicitly interpreted for this example. If 2 atm of ammonia comes to equilibrium
with its decomposition products, the equilibrium state will be 1.9572 atm of NH 3, together with
0.0214 atm of N2 and 0.0642 atm of H2. Slight further decomposition, at these concentrations,
would cause no change in free energy; there is no tendency for further decomposition because the
system is now at equilibrium. An equilibrium constant for a reaction involving gases, in which
pressures appear as an approximation for activities, is often written K P (and, because pressures
are substituted for dimensionless activities, K P may have apparent dimensions of powers of
pressure).
Consider now the problem of finding the equilibrium pressure of oxygen above Ag 2O at
o
25 C.

Ag2O ------6 2 Ag + O2

(a ) (a )
2 1/ 2

K=
Ag O2

a Ag2O
Go = 10,840 = RT ln K
K = 0.0125= PO2 ( ) 1/ 2

PO2 = 1.56 x 10-4 atm = 0.119torr

Thus silver oxide, placed in an evacuated system, should decompose until the oxygen pressure
reaches 0.119 torr. In an atmosphere containing 20% oxygen the reaction would be driven in the
other direction, so that silver should react with oxygen to form silver oxide. The slow rate of this
reaction (in either direction) is attested by the durability of silver goods. An equilibrium constant
for dissociation of a compound is often written K diss or, when the dissociation is into ions, Kionization.
For the solubility of CaF2 in water at 25oC,

CaF2 ------6 Ca++ + 2 F-

(a )(a )
Ca+ + F
2

K=
aCaF2
Go = 55900= RT ln K
K = 1.78 x 10-10

The equilibrium constant for dissolution of a slightly soluble salt in water is called the solubility
product constant, or Ksp.
The activity of solid CaF2 is 1. Assume that activities of the ions are numerically equal to
concentrations. From the dissociation equation we find that if the concentration of Ca ++ is s, the
concentration of F- will be 2s, so long as there is no other source of these ions than CaF 2.
Inserting these symbols into the equation above gives

7/10/07 4- 92
1.78 x 10-10 = s(2s)2
s = 3.5 x 10-4 M

This is the solubility of CaF2, because each mole of CaF2 going into solution gives one mole of
Ca++ ions. Introduction of activity coefficients would probably change the calculated value by
something on the order of 20%, which is often negligible in determining solubilities. (Reported
experimental values differ by much more than this for some compounds, but such accuracies are
entirely adequate for some applications.)
To find the solubility of CaF2 in 0.5 M NaCl the deviation of activity coefficients from 1
should not be neglected. The equations may be written

K = 1.78 x 10-10 = (aCa++)(aF-)2 = (Ca++)(cCa ++)(F -)2 (cF -)2


= Q ()Q (c)

Experimental measurements indicate that Q (), the correction factor, is on the order of 0.1 for
such a solution. Solving for the concentrations as before gives

1 . 78 x 10 -10
4s =3

0 .1
s = 7 . 6 x 10 - 4 M
The increase in solubility when the other ions (Na + and Cl -) are added is called the salt effect. The
added salt lowers the activity coefficients of all ions present.
The solubility of CaF2 in 0.5 M NaF will be smaller because of the common-ion effect. This
is apparent when we substitute numbers into the equations obtained above.

K = 1.78 x 10-10 = Q ()(cCa++)(cF-)2

Now the concentration of fluoride is no longer equal to twice the concentration of calcium ion,
but is 0.5 + 2s when the Ca++ has the concentration s. The correction term may be assumed the
same as for NaCl. Then

1.78 x 10-10 = 0.1s(0.5 + 2s)2

Assume first that 2s is negligible with respect to 0.5. Then

1 .78 x 10 -10
s=
0 .1 x 0.5
= 3 .5 x 10 -9 M
This answer confirms the assumption that taking s negligible with respect to 0.5 gives a self-
consistent, and therefore mathematically valid, solution to the problem.
An orderly procedure in working problems of chemical equilibrium will greatly decrease the

7/10/07 4- 93
opportunity for mistakes. The recommended steps are as follows:
1. Write the chemical reaction, even though it is a familiar one.
2. Write out the equilibrium constant expression, Q (a), in terms of activities and set this
equal to Keq.
3. Determine the equilibrium constant for the reaction as written, either from the constant
given or from standard-state free energies. Note that reversing a reaction inverts the equilibrium
constant, and that multiplication of the equation by an integer raises the equilibrium constant to
that power.
4. Make the appropriate substitutions of symbols (that is: concentrations, pressures, mole
fractions, or unity) for each of the activities. Include activity coefficient if necessary (either
numerically or symbolically).
5. Determine, from information given and the stoichiometry of the chemical reaction, what is
known about concentrations or pressures.
6. Solve the algebraic equation.
7. Examine the solution to be sure that the answer found is the answer to the question asked
and is a reasonable value in terms of the physical problem as stated and the approximations made
in the solution of the problem.

Consider, for example, the decomposition of the gas phosphine, PH 3 (G o = 18.24 kJ/mol at
o
25 C), to give white phosphorus (solid) and hydrogen gas. If 5 atm of PH 3 in a closed vessel
comes to equilibrium with its decomposition products, what will be the final pressure of PH 3?
The specific steps of the solution, as described above, are:

1. 2 PH3W 2 P + 3 H2
a P2 a H3 2
2. K eq = 2
a PH 3

3. Go = - 36.5 kJ = - 8.314 x 298 ln K


K = 2.51 x 106

6
a H3 2
4. aP = 1; 2.51 x 10 = 2
a PH 3

5. 2 PH3 -----6 2 P + 3 H2

5-2x 3x x = pressure, atm

2.51 x 106 = (3x)3/(5 - 2x)2

6. Find the solution to the equation 2.51 x 106(5 - 2x)2 = 27x3.

Because Keq is large, the reaction must be shifted well to the right, so x will be nearly equal to 5/2,
and 5 - 2x will be very sensitive to x. Therefore, as a first approximation, let x = 5/2 on the right-
hand side and find (5 - 2x):

7/10/07 4- 94
(5 - 2x)2 = 27(5/2)3/2.51 x 106 = 1.68 x 10-4
x = 2.49

This calculated value is in excellent agreement with the value assumed, so the calculation is self-
consistent.
7. The final pressure of PH3 is 5 - 2x = 0.013 atm and of H2, 3x = 7.48 atm. This is a
reasonable answer for the problem given, because G o(PH 3) = 18.24 J/mol indicates that PH3 is
quite unstable.

Temperature Dependence of Equilibrium Constants


The equilibrium constant does vary with temperature. From equation 17, ln K = - G o/RT,
so we can find the temperature dependence. Take the derivative with respect to temperature
(holding the pressure constant at its standard value), employing equation 30 of Chapter 2.

d ln K
=
d G o 1 d G o
=
( )
G o d T 1 ( )
dT dT RT RT dT R dT
S o G o TS o + G o
= + =
RT RT 2 RT 2

and therefore
d ln K H o
= (18)
dT RT 2

The last equation can be integrated, assuming H o is constant with temperature, to give

K2 H o T
ln = (19)
K1 RT1T2

In these equations K2 is the equilibrium constant at the temperature T2 , K1 the equilibrium


constant at T1, Ho is the enthalpy change for the reaction with all reactants and products in their
standard states, and T = T2-T1.
The similarity of equation 19 to the Clausius-Clapeyron equation (equation 6, Chapter 3) is
not accidental. Equations for solubility, for reaction rate constants, and for other quantities, have
equivalent forms. This form is to be expected because all of these physical properties depend on
the system surmounting an energy barrier (corresponding to the heat of reaction, the heat of
vaporization, the heat of solution, or some other such energy, or enthalpy, term) and therefore
they depend, in a fundamental way, on the Boltzmann distribution law, which is an exponential
expression relating numbers of molecules, energy states, and temperature, and which therefore
gives rise to a logarithmic temperature dependence.

7/10/07 4- 95
Electrochemistry
The best method of directly measuring the free-energy change of a chemical reaction is often
to carry out the reaction in an electrochemical cell. The free-energy change, at constant
temperature and pressure, is (eqn. 26, Chapter 2)

(T,P) dG = wrev + PdV = wrev


or
G = Wrev (20)
where Wrev is the electrical work. Electrical power, P, is I 2R = EI and electrical work is power
multiplied by time, or potential multiplied by charge:

Wrev = Pt = EI t = Eq

The potential, E, is in volts, the current I in amperes, the time t in seconds, the resistance, R, in
ohms, the charge q, in coulombs, and power, P, in watts, and the work w in joules. A convenient
unit of charge is a mole of electrons (6 x 10 23 electrons), called a faraday. One faraday, F, is
96,487 coulomb. If n mol of electrons (or n equivalents) are transferred at a potential
difference E the work done is Wrev = Eq = E(nF) and the free-energy change is

G = - nFE (21)

Thus G is obtained directly from E.


For example, we will find that the standard cell potential (all reactants and products present
at unit activity) for the reaction that transfers 2 electrons from copper to ferric iron (Fe +++) is
1.100 V. The standard free energy (per mol Cu 6 Cu++) is therefore
Go = - nFE = - 2 x 96,487 x 1.100 = - 21,230 J/mol
Substitution of equation 21 into equation 7 gives

-nFE = - nFEo + RT ln Q(a)

and therefore

ln Q (a )
RT
E = Eo (22)
nF

This is now commonly called the Nernst equation. Similarly, by combining equation 21 with
equation 17 we obtain

RT
E o
= ln K (23)
nF

7/10/07 4- 96
Writing
ai = i ci

for each of the reactants and products, the function Q factors to give

Q(a) = Q()Q(c)

and thus

ln Q ( ) ln Q (c )
RT RT
E = Eo
nF nF

If each of the activity coefficients, i , is unity, then Q() = 1, ln Q() = 0, and

( i = 1) E = Eo
RT
lnQ (c ) (24)
nF

The potential depends on the actual activities, or concentrations, of the reactants and products.
Not only G and equilibrium constants can be determined electrochemically, but also H and
S. From equation 21 and equation 30 of Chapter 2,

G E
S = = nF (25)
T P T P
and therefore, from H = G + TS, we obtain
E
H = nFE + nFT (26)
T P

HALF-REACTION POTENTIALS. An electrical potential, or electrochemical potential, exists for any


chemical reaction in which electrons are transferred from one substance to another. Such
reactions are called oxidation-reduction reactions, or simply redox reactions. Removal of
electrons, as in Fe --6 Fe+3 + 3e, or 2 I- --6 I 2 + 2e, is called oxidation; addition of electrons is
called reduction.
Electrons enter or leave a device through a conductor called an electrode. The cathode
(from the Greek down way) is the electrode at which electrons enter any device; the anode
(from the Greek up way) is the electrode at which electrons leave any device. To push
electrons through a vacuum tube or electroplating cell, the cathode must be made negative. On
the other hand, the electrode at which an electrochemical cell releases electrons is called negative
(Figure 1) and this is, by definition, the anode.
It is not possible to measure half a reaction. Only complete reactions (cathode reaction +
anode reaction) can be measured. Nevertheless, it is convenient to consider the electrochemical

7/10/07 4- 97
potential of a reaction as a sum of the potentials of
two half reactions, each defined from equation 21
in terms of the free-electron changes. Let G A be
the free-energy change for the anode reaction
(electrons leaving, hence oxidation), G C be the
free-energy change for the cathode reaction
(electrons entering, hence reduction), and G be
the free-energy change for the total reaction.
Free-energy changes are additive, so

G = - nFE = GA + GC = - nAFEA - nCFEC

and because nA = nC = n for any balanced reaction,

E = EA + E C (27)

Note that the potentials are independent of n and


thus independent of the amount of reaction! That
is, the potential of a half-reaction such as

M++ + e -----6 M

is identically the same as the potential of the half-


reaction
M++ + 2 e -----6 M

In nearly all electrochemical cells there is a


small additional term that should be included on
the right-hand side of equation 27, arising from the
contact between dissimilar solutions. Like E A and
EC , these junction potentials cannot be
measured, nor can they be calculated by
thermodynamics. They can, however, be

Because of the inherent asymmetry of an


electrochemical cell, oxidation, occurring at the
anode, with release of electrons, and reduction,
occurring at the cathode, with removal of electrons
from the circuit, produces a potential difference
between anode and cathode that drives charges
around the circuit. Charge is conserved. There is
no source of charge, or source of current.

estimated, based on theories involving diffusion rates, and are usually sufficiently small in practice

7/10/07 4- 98
that they can be ignored for present purposes.
Although it is not possible to measure EA or EC separately, it is possible to measure total cell
potentials. Then, by choosing an arbitrary value for some half-cell potential as a reference level,
other half-reaction potentials can be determined relative to this arbitrary reference. The half-
reaction chosen as the reference for this purpose is

2 e + 2 H+ -----6 H2

for which Eo (the potential with all reactants and products at unit activity) is arbitrarily assigned
the value zero. Then, for example, because the standard potential of the cell reaction

Zn + 2 H+ -----6 H2 + Zn++

Table 1 STANDARD ELECTRODE POTENTIALS, 25 OC (STANDARD REDUCTION POTENTIALS)


Reaction Eo (volt) Reaction Eo (volt)
Li+ + e ----6 Li -3.0401 Cu + e ----6 Cu+
++
0.153
K+ + e ----6 K -2.931 SO4= + 4 H+ + 2e ----6
Na+ + e ----6 Na -2.71 H2SO3 + H2O 0.172
Mg++ + 2e ----6 Mg -2.372 AgCl + e ----6 Ag + Cl- 0.22233
Al+++ + 3e ----6 Al -1.662 Cu++ + 2e ----6 Cu 0.3419
Mn(OH)2 + 2e ----6 Cu+ + e ----6 Cu 0.521
Mn + 2 OH- -1.5 I2 + 2e ----6 2 I- 0.5355
Cr++ + 2e ----6 Cr -0.913 MnO4- + e ----6 MnO4= 0.558
Zn++ + 2e ----6 Zn -0.7628 MnO4- + 2 H2O + 3e
2 CO2(g) + 2 H+ + 2 e ----6 MnO2 + 4 OH- 0.595
----6 H2C2O4(aq) -0.49 O2 + 2 H+ + 2e ----6 H2O2 0.695
Fe++ + 2e ----6 Fe -0.447 Fe+++ + e ----6 Fe++ 0.771
Cr+++ + 2e ----6 Cr++ -0.407 Hg2++ + 2e ----6 2 Hg 0.7973
Cd++ + 2e ----6 Cd -0.4030 Ag+ + e ----6 Ag 0.7996
AgI + e ----6 Ag + I- -0.15224 Hg++ + 2e ----6 Hg2++ 0.851
Sn++ + 2e ----6 Sn -0.1375 2 Hg++ + 2e ----6 Hg2++ 0.920
Pb++ + 2e ----6 Pb -0.1262 NO3- + 4 H+ + 3e ----6
O2 + H2O + 2e NO + 2 H2O 0.957
----6 HO2- + OH- -0.076 Cr2O7= + 14 H+ + 6e
Cu(NH3)4++ + 2e ----6 2 Cr+++ + 7 H2O 1.350
----6 Cu + 4 NH3 -0.05 Cl2 + 2e ----6 2 Cl- 1.35827
Fe+++ + 3e ----6 Fe -0.037 MnO4- + 8 H+ + 5e ----6
2 H+ + 2e ----6 H2 0.0000 Mn++ + 4 H2O 1.507
AgBr + e ----6 Ag + Br- 0.07133 Cr++++ + e ----6 Ce+++ 1.72
Hg2Br2 + 2e ----6 MnO4= + 4 H+ + 3e ----6
2 Hg + 2 Br- 0.13923 ----6 MnO2 + 2 H2O 1.679
Sn++++ + 2 e ----6 Sn++ 0.151 H2O2 + 2 H+ + 2e ----6 2 H2O 1.776

has been found to be 0.763 V, it follows that the standard potential for the half reaction

Zn -----6 Zn++ + 2 e

7/10/07 4- 99
must be 0.763 V. In turn, from the measured Eo for the reaction

Zn + Cu++ -----6 Cu + Zn++

of 1.100 V, it follows that for

2 e + Cu++ -----6 Cu

the standard potential is Eo = 0.337 V. We can proceed in this fashion to find the potential for
any half reaction on this arbitrary scale. Reversal of a cell reaction changes the sign of a
potential.4 The more positive a half-reaction potential, the greater the tendency of that half
reaction to proceed as written. (But it should be kept in mind that the reaction may be too slow
to observe or that some other spontaneous reaction may occur instead.)
Some standard potentials for reduction half-reactions are given in Table 1. According to the
Stockholm Convention, such standard reduction potentials are called standard electrode
potentials. If the reaction is written in reverse order, E o will have the opposite sign and is called
the standard oxidation potential. Tabulations of half-reaction potentials are often given for
oxidation half-reactions.5

CALCULATIONS OF CELL POTENTIALS. A few examples will illustrate the applications of standard
potentials of half reactions. Consider the cell reaction

2 Fe+++ + Cu -----6 Cu++ + 2 Fe++

This consists of the two half reactions,

2 e + 2 Fe+++ -----6 2 Fe++


Cu -----6 Cu++ + 2 e

The standard potential for the first is taken directly from Table 1, that for the second is obtained
from the table simply by changing the sign (because the reaction is the reverse of that given in the

4
An electric potential difference is the work required, per unit of charge, to move a charge
between two points, just as a gravitational potential difference is the work required, per unit of
mass, to move a mass between two points. The gravitational potential difference for a well is
positive if the top is measured with respect to the bottom, or negative if the bottom is measured
with respect to the top, and an electric potential that is positive if the cathode is measured with
respect to the anode will be negative if the reaction is reversed. (The standard electrode
potential includes a choice of sign in the definition, but standard potential simply means that
the reactants and products are in their standard states.)
5
For more half-reaction potentials see especially W. Latimer, Oxidation Potentials, 2nd
ed., Prentice-Hall, Englewood Cliffs, N.J., 1952.

7/10/07 4- 100
table). Thus

Eo = 0.771 + (-0.337) = 0.434 V

The positive cell potential indicates that the original reaction will proceed as written, with
substances at unit activity.
The cell reaction shown in Figure 2 is

Zn 6 Zn++ + 2e and Cu++ + 2e 6 Cu


The potential, at 25oC, may be found as follows. From Table 1,
Eo = 0.763 + 0.337 = 1.100 V

RT a Cu a Zn + +
E = Eo ln
nF a Zn a Cu + +
8 . 314 J/mol K x 298 K 0 . 01
= 1 . 130 V ln
2 equiv/mol x 96,487 C/equiv 0 .1
= 1 . 130 V
assuming the solid metals, Zn and Cu, are at unit activity (pure
and not seriously strained). The activities of the ions have been
approximated by their concentrations.
The concentration correction factor, (RT/nF) ln Q (a),
appears sufficiently often that it is helpful to carry out a partial
evaluation. At room temperature, 25oC, we may write

ln Q(a ) = x 2.306 log Q (a )


RT 8.314 x 298.15 V
nF n x 96487

ln Q(a ) = log Q(a )


RT 0.05915
(25 o C) (28)
nF n

Equation 22 may therefore be written

(25 C)o
E = Eo
0.05915
n
logQ (a) (28a)

The Nernst equation may also be applied to half-reaction potentials. For example, the
potential of the reduction half-reaction

Zn++ + 2 e -----6 Zn

would be

7/10/07 4- 101
RT a
E = Eo ln Zn
nF a Zn + +

or, at 25oC,

0 .05915 a
E = 0 .763 log Zn
2 a Zn + +

(The electrons should, in principle, also appear in this expression, but the activity of the
electrons is unknown, and when this reduction half-reaction is combined with an oxidation half-
reaction, the electron terms will drop out.) Similarly, for the oxidation half-reaction
Zn -----6 Zn++ + 2 e
the potential, at 25oC, would be

0 . 05915 a ++
E = 0 . 763 log Zn
2 a Zn

Although space does not permit an adequate discussion of the problem of reversibility, it
should be pointed out that not all chemical reactions can be carried out in electrochemical cells,
and even many that can will not be reversed by a change of potential. The first requirement for a
well-behaved electrochemical cell must be that the reaction will proceed as desired and be capable
of reversal. To achieve quantitative agreement with the equations of thermodynamics, however,
the requirements are more stringent. The chemical reaction must be thermodynamically
reversible, so that an infinitesimal change of potential is sufficient to reverse the flow of current
and the direction of chemical reaction. Discussions of the problems encountered in irreversible
electrochemical systems may be found in more advanced works under the listings of
overvoltage, polarization, and junction potentials.

CELL NOTATION. A very convenient shorthand notation has been developed for writing
electrochemical cells. The cell shown in Figure 2 can be represented by
Zn/ZnSO4(0.01 M)//CuSO4 (0.1 M)/Cu
A cell in which the reaction is the displacement of hydrogen by zinc can be written
Zn/Zn++//H+/H2,Pt
The platinum serves as the electrical contact to the solution, with hydrogen gas bubbled around
the platinum. Although there are minor differences among authors concerning separating
components with commas and shilling marks, generally one of these marks will separate
substances in direct contact. The // indicates a junction between two different solutions, which
are commonly connected by a salt bridge, such as a U tube containing concentrated KCl

7/10/07 4- 102
solution and dipping into the two solutions.
The direction of the cell reaction is determined by the convention that the electrode on the
left is the anode.6

APPLICATIONS OF ELECTRODE POTENTIALS. Electrochemistry has important applications in


physical chemistry, analytical chemistry, descriptive inorganic chemistry, preparative inorganic
and organic chemistry, biochemistry, and metallurgy. Electrochemical measurements provide an
accurate means of obtaining the thermodynamic properties of a reaction, including H, S, and
G. These thermodynamic properties depend, in turn, on the exact nature of the reaction,
including the charges on the ions participating. Because the potentials depend on concentrations,
information is available on the concentrations of individual ions in solution for analytical purposes
or for determining solubilities.
Electrode potentials are often more convenient than free energies for prediction of which
reactions will be spontaneous. For example, oxidizing agents can be put in order of oxidizing
powers by reference to the standard potentials for the corresponding reactions. Often an
oxidizing agent can be found that is just strong enough to do the job desired and not so strong
that it will oxidize other substances present in the mixture. Electrode potentials are the basis of
studies of corrosion and electro-refining of metals.

Problems
1. The standard free energies of NO2 and N2O4 at 25oC are 51.3 and 99.8 J/mol. Find the
equilibrium constant for the reaction

2 NO2 ----- N2O4

2. For each of the following reactions, set up the equilibrium constant expression, in terms of
activities, and make the appropriate substitutions of symbols (concentrations, pressures, mole
fractions, or one) for the activities:
a. Cu++ + H2S(aq) ----- CuS + 2 H+
b. 2 HgO ----- 2 Hg + O2
c. NaHCO3 + HCl ----- CO2 (g) + H2 O + NaCl
H 2O

d. AcOEt + H2O ----- AcOH + EtOH


(ethyl acetate) (acetic acid) (ethanol)

6
There are very few conventions that need to be memorized, although these few can be
expressed in a bewildering variety of forms. One convention is that the anode (where the
electrons leave the cell) in on the left in these condensed cell descriptions. A mnemonic for
remembering this is a partial acrostic based on the word always: Anode ... Left ...W ...A ...Y
...S. Another convention is that the term standard electrode potential means the same as
standard reduction potential. A third is that a positive potential for a cell corresponds to a
spontaneous reaction, as implied by equation 2.

7/10/07 4- 103
3. The standard free energies of benzene and cyclohexane are 124.5 and 28.5 J/mol at 25 oC. Find
the equilibrium constant for the reduction of benzene with hydrogen at 25 oC.

C6H6 (liq) + 3 H2 ----- C6 H12 (liq)

If equilibrium is reached with 5 atm H 2 gas present, what is the mole fraction of benzene left in the
cyclohexane liquid produced?
4. What effect will the addition of an inert gas have on the equilibrium point of the gas-phase
reaction
N2O4 ----- 2 NO2
if the volume is maintained constant?
5. What effect will the addition of an inert gas have on the equilibrium point of the gas-phase
reaction
N2O4 ----- 2 NO2
if the total pressure is maintained constant?
6. For the reaction

CO + 2 H2 ----- CH3OH

Go = -13.5 J at 700 K. Find the percentage decomposition of methanol, at this temperature and
constant volume, if the initial pressure of pure methanol is 5 atm.
7. Water and solid sulfur can react to form SO 2 and H2S according to the equation

H2O (gas) + 3/2 S ----- SO2 + H2 S

The standard free energies are, for H 2O (gas), -228.6, for H2S (gas) -33.4, and for SO2 (gas), -
300.1 J/mol, at 25oC.
a. Find the equilibrium constant for the reaction as written.
b. Calculate the pressures of SO2 and H2S to be expected in equilibrium with solid sulfur and
moist air (vapor pressure of water = 23.8 torr) at 25 oC, assuming gases to be ideal.
8. The reaction
2 SO2 + O2 ----- 2 SO3

is of considerable industrial importance. Starting with one-fourth atm of SO 3 vapor at 25 oC, in a


fixed volume, find the equilibrium pressure of SO 3 and hence the percentage dissociation.
9. The standard free energy of HI at 250oC is -10.8 J/mol.
a. Find Go and the equilibrium constant for the reaction at 250o C,

H2 + I2 ----- HI
b. Find Go and the equilibrium constant for the reaction, at 250o C,

H2 + I2 ----- 2 HI

7/10/07 4- 104
How does this K compare with the K from part a?
c. If 1 atm of HI at 250oC is placed in a vessel and allowed to reach equilibrium with its
decomposition products at this temperature, what will be the final pressure of HI?
10. The solubility product constant, Ksp , for MgF2 at room temperature is 7.1 x 10 -9. Assuming
activity coefficients are unity, find the solubility of MgF 2
a. in water
b. in 0.01 M NaF solution
Addition of ions to a dilute solution will decrease the values of activity coefficients. Would the
addition of Na2SO4 increase or decrease the solubility of MgF2 ?
11. Find Go for the process of dissolving MgF2 in water. Is this positive or negative? Explain
why, in terms of what will happen when MgF2 is added to water.
12. The solubility products for barium oxalate at 18 oC are given below.

(A) BaC2O42 H2O Ksp = 1.2 x 10-7


(B) BaC2O4 H2O Ksp = 2.2 x 10-7

a. Assuming only (A) is present in the solid phase, find the concentration of Ba ++ ion in
aqueous solution in equilibrium with this solid.
b. Assuming only (B) is present in the solid phase, find the concentration of Ba ++ ion in
aqueous solution in equilibrium with this solid.
c. Describe what will happen if 1 g of (A) and 1 g of (B) are added to 100 ml of water at
o
18 C under 1 atm pressure.
d. How many phases will be present at equilibrium? What are they?
13. The dissociation constant of H2S in water solution is 9.1 x 10 -8.

H2S ----- H+ + HS-

(The second dissociation, to give S=, can be ignored here.) Find the concentration of hydrogen
ion in 0.1 M H2S solution.
14. The dissociation constant of benzoic acid is 6.30 x 10 -5 at 25 oC. Find the concentration of
hydrogen ion
a. in 0.001 M benzoic acid solution
b. in solution containing 0.001 M benzoic acid plus 0.01 M sodium benzoate. (Benzoic acid
dissociates to give one hydrogen ion, BH ----- B- + H+.)
15. For the cell
Zn/Zn++(a = 0.01)//Ag+(a = 0.5)/Ag

a. Write the over-all cell reaction.


b. What is E o?
c. What is E ?
d. Is the reaction spontaneous?
16. Find the equilibrium constant for the reaction, at 25 oC,

7/10/07 4- 105
2 Fe++ + Hg2++ ----- 2 Hg + 2 Fe+++

17. The potential of the cell

H2(1 atm)/HCl(0.01 M),AgCl(s)/Ag(s)

is 0.464 V at 25oC.
a. Will the potential be increased or decreased by adding NaCl to a concentration of 0.10 M?
b. Will the potential be increased or decreased by alloying the silver with gold?
c. Is the silver the anode or cathode?
d. As the cell is written, will electrons in the external circuit flow from left to right or from
right to left?
18. The Weston normal cell is Cd-Hg/CdSO4/Hg2SO4/Hg, for which the potential is E = 1.0183
[1 - 4.06 x 10-5(t - 20) - 9.5 x 10-7(t - 20)2 + 1 x 10-8 (t - 20)3 ] V at to C. Find H for this cell at
20oC.
19. The cadmium-calomel cell

Cd + Hg2Cl2 ----- Cd++ + 2 Cl- + 2 Hg

has been found to have a potential that varies with temperature according to the equation

E o = 0.6708 - 1.02 x 10-4(T - 298) - 2.4 x 10-6(T - 298)2


with T in kelvin and E o in V. For the cell reaction at 40o C, calculate
a. Go b. Ho c. So
20. For the cell
Pt, H2(1 atm)/HCl/AgCl/Ag
in which the electrolyte also contained 0.001 M NaCl, the measured potential, at 25 oC, was found
to be 0.724 V. What is the pH of the solution, assuming activities equal to concentrations?
21. Write the reaction to be expected if a block of Mg is attached to a steel ship hull. Will the
ship be oxidized or protected from oxidation?
22. Find an oxidizing agent capable of oxidizing Hg to Hg2++ without appreciable formation of
Hg++.
23. Predict the products of the following reaction:
a. HNO3 + Hg -----
b. HNO3 + Hg2++ -----
c. Hg + Hg++ -----
24. In deriving equation 27 it was emphasized that electrochemical potentials are independent of
the amount of reaction. (Holding down the button on a flashlight does not increase voltage.) Yet
the Nernst equation (see equation 28a) contains n explicitly. How can you reconcile these
seemingly contradictory facts?
[Answer: 24. Q(a) contains powers of activities proportional to n. Dividing log Q(a) by n
removes this apparent dependence on n from the correction term for E. ]

7/10/07 4- 106
5 Heat Engines and Absolute Zero
Many innovations of thought and technology contributed to the Industrial Revolution, but
certainly a major factor in its development was the replacement of human and animal power by
the power of the steam engine and, later, by power from the internal combustion engine.
Thermodynamics, too, owes much of its early development to these engines that convert thermal
energy to mechanical energy (heat to work). It was through his studies of heat engines in the
early 19th century that the young Frenchman Sadi Carnot discovered the principle now best
described in terms of entropy, and even today the heat engine is considered important as a means
of demonstrating the principles of classical thermodynamics. But the principles of the heat engine
are also of great practical importance in modern technology and in current research. New devices
are constantly being sought to convert thermal energy into electrical power. One way of stating
the second law is that mechanical energy (as work) can be totally converted to thermal energy (or
heat) but thermal energy cannot be totally converted to work (i.e., mechanical energy) in a
cyclic process. Thermodynamics provides an understanding of the possible effectiveness, and the
inherent limitations, of all devices to convert thermal energy into mechanical energy. The theory
of heat engines is also important to a discussion of absolute zero and the third law of
thermodynamics.

Heat Engines and Heat Pumps


The essential ingredients of a steam engine are a boiler and a source of heat for it; a cylinder
into which the steam from the boiler can expand fitted with a piston against which the gas
will push, performing work (and thereby losing energy and becoming cooled); and means for
further cooling and/or venting the expanded steam so that the piston can be returned to its initial
position. In the operation of the engine, heat is put into the engine at some elevated temperature,
the engine itself does work on its surroundings (transmitted by the shaft of the piston), and a
portion of the thermal energy taken in is discharged, as heat, at a lower temperature (into the
cooling water). The energy transfers are represented in Figure 1.

Figure 5.1. A heat engine takes thermal energy from a reservoir at a higher temperature,
TH, does work on the surroundings, and dumps excess thermal energy at TC .

7/10/07 5- 107
Theoretical heat engines always operate on a completely reversible, closed cycle. Real heat
engines seldom, if ever, operate on a fully closed cycle. Expanded steam, for example, is typically
at least partially released to the atmosphere, rather than being condensed to liquid and recycled.
That has little effect on the overall efficiency, so we retain the cyclic model.
The efficiency of a heat engine is defined to be the output, (-) W, divided by the input, Q U,
which is the thermal energy put into the engine at the upper temperature, T U .
output - W
efficiency= = (1)
input QU

For a complete cycle, QU + QL + W = E = 0, and QU is positive, QL, which is the thermal energy
(or heat) absorbed by the engine at the lower temperature, is negative, and W, the work done on
the heat engine, is negative. Thus, in terms of all positive values
(-W) = QU - (- QL)
or
|W| = |QU | - |QL |

Therefore QU > (- W) and the efficiency is less than one (or less than 100%).

CARNOTS THEOREM. Between 1698 and 1782, a series of British inventors designed steam
engines, originally to remove water that flooded the coal mines but eventually to initiate the
Industrial Revolution that was largely responsible for the creation of the British Empire. Toward
the end of the century, Lazare Carnot, a French mathematician, engineer, general, and
administrator under Napoleon, studied the design of steam engines and concluded that efficiency
was decreased by any jerkiness (sudden acceleration) of the engine. His son, Sadi, followed in
his footsteps as an officer, until his career was terminated by Napoleons collapse. Sadi then
retired, at half pay, at the age of 24 (in 1820) to study physics, economics, and steam engines. In
1824 he wrote a small pamphlet, Reflections on the Motive Power of Fire, that laid out the theory
of heat engines, including a derivation of the maximum possible efficiency of any such device.
Consider two heat engines, A and B, operating between the same two temperatures. We
wish to avoid any jerkiness in their operation, so we may imagine that engine B, at least, operates
reversibly. We put no restrictions on engine A, so we may start with the trial assumption that the
efficiency of A is greater than that of B. The two engines can then be coupled together as shown
in Figure 2. Engine A drives engine B in reverse, and because (-W)/Q U is greater for engine A, we
set (QU)A = - (QU)B (by adjusting the length of the stroke). Then (W)B > (-W)A , so there is more
than enough mechanical energy available to drive engine B. By allowing some input of energy at
the lower temperature, from the cooling water, we now have a pair of engines that will give us
perpetual motion (of the second kind). Engine A drives engine B with excess work available to
carry out some other task. But we know this is impossible, both from long experience and from
our understanding of the second law of thermodynamics.
The conclusion of this thought experiment is that the efficiency of engine A cannot exceed the
efficiency of any reversible heat engine, such as B. This conclusion, reached by Sadi Carnot, is
known as Carnots theorem:

7/10/07 5- 108
Figure 5.2. Two heat engines may be coupled, such that the reversible engine (left)
provides thermal energy at the higher temperature to the other engine, which in turn
provides the work to drive the first engine.

The efficiencies of any two reversible heat engines, operating between the same
two temperatures, are the same, and this is the maximum efficiency
for any heat engine operating between these two temperatures.
From Carnots theorem we can see that the nature of the working fluid whether steam,
ideal gas, mercury, or any other vapor, liquid, or vapor-liquid combination cannot change the
efficiency, nor can the nature of the cyclic process alter the efficiency, so long as the steps are
reversible and thermal energy transfers occur only at the upper and lower temperatures. The
efficiency will, in fact, be decreased by non-reversible heat transfers or transfers at intermediate
temperatures and, to a lesser extent, by friction, as we will see below, but these are differences
associated with the departure from reversibility.
It is sufficient, therefore, to choose a very simple cyclic process, operating with an ideal gas.
The result will be applicable to any real device, apart from corrections for irreversibility.

CARNOTS CYCLE. The cycle most conveniently treated is called the Carnot cycle. It consists
of an isothermal reversible expansion at the upper temperature, an adiabatic reversible expansion
from the upper temperature to the lower temperature, an isothermal reversible compression at the
lower temperature, and an adiabatic reversible compression from the lower to the upper
temperature. More briefly, it alternates isothermal and adiabatic steps, with every step reversible.
The important equations of the Carnot cycle are obtained quite easily for an ideal gas.
Step 1: at T1, E1 = 0; V1 6 V2:
V2
W1 = Q1 = nRT1 ln (2a)
V1

Step 2: Adiabatic, Qa = 0; T1 6 T2; V2 6 V3:

Wa = Ea = n CV (T2 - T1) (2b)


and in more detail,
dEa = nCV dT= wa = - PdV= - nRT dV/V so CV ln T2 /T1 = - R ln V3 /V2

7/10/07 5- 109
Step 3: at T2, E2 = 0; V3 6 V4:
V4
W2 = Q2 = nRT2 ln (2c)
V3

Step 4: Adiabatic, Qb = 0; T2 6 T1; V4 6 V1:

Wb = Eb = n CV (T1 - T2) (2d)


and in more detail,
dEb = nCV dT= wb = - PdV= - nRT dV/V so CV ln T1 /T2 = - R ln V1 /V4
Then, because ln T1/T2 = - ln T2/T1, it follows that ln V3 /V2 = - ln V1 /V4 and thus
V2 /V1 = V3/V4, so
Q1 Q
= 2 (3)
T1 T2

and

= nR (T 2 T 1 ) ln
V2
W = W i
V1
The efficiency is
output W T1 T2 T
= = = (4)
input Q1 T1 T1

SHORTCUT: We may ignore the details of the ideal-gas expansions and compressions and, in fact,
remove the restriction to an ideal gas, by emphasizing the changes in entropy.
In the initial reversible isothermal step at T 1, E 1 = 0 and

- W1 = Q1 = T1 S1

For the reversible adiabatic step (T 1 to T2), Q is zero and therefore S is zero. In the reversible
isothermal step at T2, E2 = 0 and

W2 = - Q2 = - T2 S2

and in the reversible adiabatic step returning from T 2 to T 1, Q is zero and S is zero.
Now we note that for the over-all process,

S = S1 + S2 = 0

because the engine has undergone a complete cycle. It follows that

7/10/07 5- 110
S1 = - S2

Also, because E = 0 for the complete cycle,

- W = Q = Q1 + Q2 = (T1 - T2) S1

Let S1 be represented by the symbol (Greek kappa). The important equations are then
Q1 = T1 (5a)
Q2 = - T2 (5b)
- W = (T1 - T2) = T (5c)

with Q1 the amount of thermal energy transferred to the engine at T1 , Q2 the amount transferred to
the engine at T2, and W the total amount of work done on the engine in a complete cycle. These
three equations must apply to any reversible heat engine. Only the value of will vary from one
engine to another, for a given choice of T1 and T2.

APPLICATIONS OF HEAT ENGINES AND HEAT PUMPS. The maximum possible efficiency for a heat
engine producing net work ( > 0, W < 0) is that of a reversible engine. From equation 4 this is
-W T
efficiency = = (4a)
Q1 T1

For example, a steam engine operating between 100 oC and 20oC has a maximum possible
efficiency of

-W T 80
eff = = = = 21 . 4 %
Q1 T1 373

Steam engines can be designed to operate with superheated steam, with a resultant improvement
in possible performance. Operating between 600 oC (a typical federal limit) and 30o C, the
maximum possible efficiency would be

570
eff = = 0 . 65 = 65 %
873

which still represents a loss of more than a third of the energy, and even this value is too
optimistic.

PRACTICAL LIMITATION OF CURZON AND AHLBORN. It has always been apparent that the Carnot
cycle puts an unrealistic demand on the system, because it assumes thermal energy transfer at T 1
and at T2. Such a transfer, without a thermal gradient, would be infinitely slow. Although such
an arrangement would give maximum efficiency, it would give zero power output because it could

7/10/07 5- 111
never finish its first cycle.
Curzon and Ahlborn have shown that optimum power output occurs when the demands of
rate of thermal energy transfer are balanced against the engine efficiency. When this limitation is
included, the actual best efficiency becomes

T1 T2 W
eff = = (6)
T1 Q1

This is a substantially lower efficiency than predicted by Carnot, but it agrees quite well with the
actual efficiencies obtained from real heat engines. (This shows that friction, usually blamed for
poor performance, is relatively inconsequential in well-designed heat engines.)
By the Curzon and Ahlborn formula, the efficiency of an engine operating between 20 oC and
100oC would drop to 1 - %(293/373) = 11%. Moving to the higher temperatures of 600o C and
30oC would give an efficiency of 1 - %(303/873) = 41%.

REVERSING A HEAT ENGINE. Our proof of Carnots theorem relied on the assumption that it is
possible to reverse a heat engine, which would mean taking in thermal energy at a low
temperature from the surroundings (e.g., the cooling water or from outdoors) and releasing high-
temperature thermal energy to the surroundings, as a consequence of our having done work on
the engine. Under these circumstances, our efficiency becomes the amount of thermal energy
transferred (at T1 or at T2, depending on our intent) divided by the amount of work we must do on
the engine.
To transfer thermal energy to the reservoir at T 1,
output - Q1 T
efficiency (warming) = = = 1 (7)
input W T

whereas the efficiency for removing thermal energy from the reservoir at T 2 would be
output Q2 T
efficiency (cooling) = = = 2 (7a)
input W T

Suppose, for example, we wish to move thermal energy from outside, where it is freezing
(0 C) to inside where it is 25o C. Then the calculated efficiency is 298/25 = 11.92 = 1192%.
o

Whereas we could buy 1 kWh of electrical energy and get 1 kWh of thermal energy put into the
room from an electric heater (which is necessarily 100% efficient), if we could put that same
power into a reversible Carnot engine we could get 11.92 kWh of thermal energy into the room,
most of it (10.92 kWh ) coming from the 0oC air outside. Is this possible?
The answer has to be yes and no. Yes, we can build reversible heat engines, which for the
application described would be called heat pumps. Yes, the heat pumps would extract thermal
energy from outside and dump large amounts of thermal energy inside. The same type of reversed
heat engine can also pull thermal energy from the inside of a freezer or refrigerator and dump it

7/10/07 5- 112
into the room. But no, for the same reasons discussed by Curzon and Ahlborn, we cannot get
very close to 1100% efficiency. Actual measured efficiencies typically range between 3 and 5 (or
300% and 500%), which is still quite good, but not as spectacular as Carnot would indicate.
Many people believe an efficiency should inherently be less than, or at most equal to,
100%. The output/input (i.e., efficiency > 1) of a heat pump or refrigerator is therefore typically
labeled as a coefficient of performance.

The Third Law of Thermodynamics


As we have seen, the first law of thermodynamics deals with the conservation of energy as it
is transformed and transferred between the system and surroundings. Although temperature is
important in measuring energy changes, it enters only as temperature differences, for which the
scale is quite arbitrary. The second law of thermodynamics predicts which of the many
conceivable transformations and transferrals of energy can actually occur, given sufficient time.
In the definition of the entropy function, an absolute temperature appears, which recurs in the
equations for phase equilibrium, in the expressions for the efficiencies of heat engines, and in
other applications of the second law. The third law of thermodynamics is comparatively recent,
having been proposed early in the 20th century. Several forms have been suggested for this
additional postulate, but all are intimately involved with the absolute temperature scale, and
particularly with the absolute zero of temperature.

DETERMINATION OF THE ABSOLUTE TEMPERATURE SCALE. It has not been possible to achieve
the absolute zero of temperature and there is good reason to believe that it cannot be reached,
although temperatures routinely achieved are now in the general range of nanokelvin, 10 -9K.
Measurements at very low temperatures are important, but identification of absolute zero is
critical to studies even at room temperature.
For example, embedded in Carnots equations is a definition of thermodynamic temperature,
measured in terms of thermal energy transfer. From equation 3,

T1 Q1
= (3a)
T2 Q 2

If we compare measurements at the freezing point of water and the boiling point of water, the
measured value of this ratio is 1.366, from which we find T 2 = 2.73 T and T1 = 3.73 T. Then if
we define T to be 100 K, we establish the kelvin scale with T 2 = 273 K and T1 = 373 K.
Alternatively, it is found preferable to set one fixed point, taken as the triple point of water at
273.16 K. The boiling point is then measured to be 100 K above the freezing point of water.

NERNSTS HEAT THEOREM. The entropy change for a chemical reaction at a temperature T 2 can
be found if the entropy change at some other temperature, T 1, is known and the heat capacities of
reactants and products are known. The procedure is analogous to Kirchhoffs law (equation 17,
Chapter 1). For a constant pressure,

7/10/07 5- 113
CP (reactants)
T1 T2 C (products
)
S2 = dT + S1 + P dT (8)
T2 T T1 T

From measurements of entropies of reaction near room temperature, combined with


measurements of heat capacities over a range of temperatures, Nernst observed that calculated
entropies of reaction seemed to go to zero as the temperature approached zero. This observation
provided the basis for a generalization known as the third law of thermodynamics.
There are several ways of stating the third law, and many of the implications are best
understood from the standpoint of quantum statistical mechanics. For present purposes the
following statement will suffice:
Every substance in its lowest energy state, at absolute zero, has the same entropy.
An equivalent statement would be
the entropy change for every process (chemical or physical change) becomes zero when the
process occurs between substances in their lowest equilibrium states at absolute zero.
We cannot demonstrate this by carrying out reactions at 0 K. In the first place it is impossible to
achieve absolute zero, and in the second place, the rates of chemical reactions and other processes
decrease so rapidly with decreasing temperature that few, if any, chemical reactions could proceed
at temperatures even near 0 K. It is necessary, therefore, to rely on calculated entropies of
reaction. It is now known that such calculated values include many apparent exceptions to the
third law, which generally prove to be the best supporting argument for the law.
From a knowledge of the structure of a substance, together with an understanding of
statistical mechanics, it is possible to predict which substances are likely to be apparent
exceptions to the third law and to calculate the exact amount of the discrepancy to be expected.
The excellent agreement obtained between theory and experiment on the exceptions is evidence
of the validity of the original postulate.
Assume, for example, that a substance has an energy state that differs only slightly in energy
(or enthalpy) from the lowest energy state, but that this upper energy state has an appreciable
entropy difference from the lowest state. Labeling the upper state with an asterisk and the lowest
state with a subscript zero, H* - Ho is positive but very small, whereas S* - So is positive and not
vanishingly small. Then, at a temperature greater than 0 K, T(S* - S o) may be greater than H* -
Ho, so the free-energy change for the transition to the upper state will be negative and the system
will exist predominantly in the upper state. Only for vanishingly small temperatures will the
enthalpy term be able to overcome the entropy term and make the lowest energy state most
stable, but at such low temperatures the time required for the change of state may be so great that
the lowest energy state is never achieved experimentally.
Entropy is a measure of disorder in a system and its evaluation depends upon a counting
operation, so the entropy differences between states are often expressible as a function of small
integers. Energies, however, may be arbitrarily close. Energy states that are very close together
are said to be degenerate or nearly degenerate. An example of near degeneracy of energy
levels is provided by carbon monoxide, CO, which can pack in a solid crystal in the perfectly
ordered arrangement CO CO CO or in disordered arrangements, such as CO CO OC CO, in
which some molecules are turned around. Although these have only slightly different energies,

7/10/07 5- 114
there are so many possible disordered states that the equilibrium entropy at temperatures slightly
above zero is1 R ln 2 = 5.76 J/molK above the ground state. This disorder remains, frozen in,
as the temperature is lowered toward zero.
Another form of degeneracy arises from the presence of small magnetic effects of the nuclei
(associated with nuclear spins). The magnetic field of the Earth is so weak that the nuclear
magnets are not well aligned with each other. This has a negligible effect on the energy, but again
a very appreciable effect on the entropy. However, the nuclei are unchanged in a chemical
reaction, so there is no entropy change in a chemical reaction arising from these nuclear magnetic
effects. The nuclear effects therefore do not cause an apparent exception to the third law.
Because the spin state does not change, it may be called a frozen metastable equilibrium state,
or a pseudoequilibrium state.
The third law may be restated, explicitly recognizing the existence of nonequilibrium states,
as follows:
Lim S = 0 (9)
T0

for any isothermal process involving only phases in internal equilibrium or involving any phase in a
pseudoequilibrium state, provided the process does not disturb this pseudoequilibrium.
The importance of the third law, for classical thermodynamics, is that it provides a natural
reference level2 from which entropies can be measured and tabulated. These absolute entropies
are available in handbooks, based on measurements of heat capacities from low temperatures. A
few such values are given in Table 1. Tabulated entropies can be added or subtracted to find
entropies of reaction, which can be combined with enthalpies of reaction to calculate free energies
of reaction, and hence equilibrium constants. Entropies can also be calculated, employing
equations of statistical mechanics, if the molecular properties are known from spectroscopic
measurements.
Table 1 STANDARD ENTROPIES, 25oC
Compound State So (J/molK) Compound State So (J/molK)
Al c 28.3 NO g 210.8
AlCl3 c 109.3 NO2 g 240.1
AlF3 c 66.5 N2O g 220.0
Al2O3 c () 50.9 N2O4 lq 209.2

1
The entropy depends on the number of possible molecular arrangements that the
equilibrium state comprises. For each CO molecule there are two possible positions, so for N
molecules, there are (2)N possibilities. The entropy is k ln 2N = Nk ln 2 = R ln 2 = 5.76 J/molK
if N is taken as Avogadros number.
2
This reference level is often called zero entropy, but that name is somewhat misleading.
No absolute value for entropy can be predicted from classical thermodynamics, and there is no
way in which an absolute value of entropy could be meaningfully employed The so-called zero
of entropy is usually not the lowest entropy state attainable, for it ignores the spin effects
mentioned above.

7/10/07 5- 115
Br2 lq 152.2 g 304.4
g 245.47 HNO3 g 266.9
C graphite 5.74 NH3 g 192.7
C diamond 2.4 Na c 51.3
CO g 197.66 NaCl c 72.1
CO2 g 213.78 NaOH c 64.5
CS2 g 237.8 O2 g 205.15
Cl2 g 223.1 O3 g 238.9
F2 g 202.8 P red 22.8
H2 g 130.68 white 41.1
HBr g 198.70 PH3 g 210.2
HCl g 186.90 S rhombic 32.05
HF g 173.78 SF6 g 291.5
HI g 206.6 SO2 g 248.2
H2O g 188.83 CBr4 g 358.1
H2O2 g 232.7 CCl4 g 309.4
H2S g 205.8 lq 214.4
I2 c 116.1 CF4 g 261.6
g 260.6 CH4 g 186.3
K c 64.7 C2H2 g 200.9
KCl c 82.6 C2H4 g 219.3
KOH c 78.9 C2H6 g 229.2
Mg c 32.7 CH3OH g 281.6
MgO c 27.0 lq 160.7
MgCl2 c 89.6 C2H5OH g 281.6
MgF2 c 57.2 lq 160.7
N2 g 191.01 C6H6 lq 173.4

ATTAINMENT OF ABSOLUTE ZERO.3 As we have seen, the most efficient refrigeration system is a
reversible heat pump. The work required to withdraw an amount of heat Q = - Q 2 from a low-
temperature reservoir at T2 is, from equation 7a, W = (T/T2 )Q2 . As T2 goes to zero, therefore,
the amount of work required to remove a given amount of thermal energy from any substance at
the temperature T2 will become infinite. It thus appears that the second law prohibits us from
reaching the absolute zero.
A more meaningful equation for our purposes is obtained by finding the work required for a
given change in temperature. Replacing q2 with CP dT, the differential form of equation 7a
becomes
CP
w= TdT (10)
T2

in which w is the amount of work required to change the temperature of a substance, with heat
capacity CP, from the temperature T2 to the temperature T2 - dT, and T is the difference in

3
See E.M. Loebl, J. Chem. Ed. 37, 361-363 (1960), and R. Fowler and E.A.
Guggenheim, Statistical Thermodynamics, Cambridge, 1939, pp. 224-227.

7/10/07 5- 116
temperature between the substance and the surroundings. If the ratio C P/T 2 becomes infinite, as
might be expected, when T2 goes to zero, the amount of work required for any given temperature
change of the substance would indeed be infinite. It has been found, experimentally, however,
that CP also goes to zero as the temperature approaches 0 K, and in fact the ratio CP /T2 becomes
zero, rather than infinite. Therefore, we cannot argue that the second law prohibits us from
reaching absolute zero.
If CP/T2 does go to zero as T2 goes to zero, as observed experimentally and predicted from
theory, then equation 8 can be applied to find entropies at the absolute zero. (The integral is not
defined unless CP/T remains finite.) Then it becomes possible to show that the third law of
thermodynamics prohibits us from reaching absolute zero.
The proof is as follows. Let A and B represent any two phases that differ in some respect but
which may be interconverted. For example, A and B might represent reactants and products for a
chemical reaction, or they might represent two different crystalline modifications of a given
chemical substance. Then the entropy of A at T 1 may be related to the entropy at absolute zero by
the equation

(T 1 ) = ( A ) dT
T1
+
o
S A S A C P
0 T
and similarly, for phase B, the entropy at any temperature T 2 is

S B (T2 ) = S Bo + C P (B )
T2 dT
0 T
For every temperature, T2, of phase B there must be a temperature, T1 , of phase A such that SA (T1 )
= SB(T2):

SA (T1 ) = SAo + CP ( A) = SB + CP (B) = SB (T2 )


T1 dT o T2 dT
0 T 0 T
Rearranging, we obtain

S Bo S Ao = CP ( A) C P (B)
T1 dT T2 dT
0 T 0 T
We may assume T1 $ T2; if this is not satisfied we need only relabel the phases A and B to make it
so.
Because the transformation from A, at T 1, to B, at T2, is isentropic (no entropy change), it
should be possible, in principle, to achieve an adiabatic, reversible transformation,

A(T1) -----6 B(T2)

We are free to choose either T1 or T2; we choose to let T2 = 0. Then either T1 > 0 or T1 = 0. If T1
> 0, then

CP ( A)
T1 dT
0 T
> 0 and SBo > SAo

7/10/07 5- 117
but this would be a violation of the third law, which says that A and B must have the same entropy
at absolute zero.
The only other possibility is that T1 = 0. That is, the transition
A(T1) -----6 B(0)
can occur only if T1 is zero; one cannot reach absolute zero from a non-zero temperature. The
same argument may be turned around. If SBo were greater than SAo, then T1 would be greater than
zero and it would be possible to reach absolute zero. If T 1 must be zero, then it follows that SA o =
SBo.
Therefore, the following statement may be considered as an alternative form for the third law
of thermodynamics:
It is impossible by any procedure, no matter how idealized,
to reduce any system to the absolute zero in a finite number of operations.

7/10/07 5- 118
Appendix
Basic Operations of Calculus

Calculus is an old term for calculations. It is now applied to the methods of


mathematics developed by Isaac Newton and by Gottfried Wilhelm Leibniz. Many problems of
physics, chemistry, and engineering require an understanding not only of the operations of
calculus but also of the justification and the limitations for these operations. Such questions are
properly treated in mathematics texts. This appendix is in no way a substitute for such a rigorous
development of calculus. It is, rather, a temporary expedient to allow the student who has not yet
reached some of these operations in his mathematics studies or has forgotten some details to
apply those particularly simple operations that are required in elementary thermodynamics.
What makes calculus different from ordinary algebra is that it looks at the limiting values of
quantities, including infinite1 numbers of quantities or steps. Provided the mathematical
expressions are well behaved, the resulting equations are no more difficult than algebra.

A.1 Functional Notation


Whenever the value of one quantity, or variable, depends on the value of some other
quantity, or variable, the first variable is said to be a function of the second. This is often written
y = f (x), read as y equals f of x, to indicate that the value of x determines the value of y. 2 For
example, the area, A, of a circle is a function of the radius, r. We write this A = f (r), where f (r)
= r2. The symbol f ( ) may be considered a mold into which the variable is placed. That is, if
f (r) = r2, then f (x) = x2, f (z) = z2 , and f (a) = a2 . If f (x) = 3x2 - 2x + 5, then f (z) = 3z2 -
2z + 5.
Note that we often build in mnemonic devices. For example, even though area may be
considered as a variable in a particular problem, and would therefore usually be represented by a
symbol near the end of the alphabet, we represent area by A. Furthermore, we let such symbols
do double duty by labeling functions in the same way. Thus we may write
A = A(r) = r2
letting A represent both the variable area and the function whose value gives the area.

1
We represent infinity by 4, but it is not a definite number. Infinity is a symbol, or
name, for any very large quantity that is larger than any number you (or someone else) may select
beforehand.
2
We are assuming at present that f (x) = y is a single-valued function of x; that is, for
each value of x, there is just one value of y = f (x). However, there are many situations where one
value of x may correspond to more than one value of f (x), or the same value of f (x) may arise
from more than one value of x.

7/10/07 A- 119
A.2 Theory of Limits
Zeno posed the problem of Achilles and the tortoise. Achilles could run ten times as fast as
the tortoise. If the tortoise was initially 100 m from Achilles, then when Achilles has run 100 m,
the tortoise has moved only 10 m. When Achilles has run 10 m, the tortoise has moved 1 m.
Each time Achilles runs the distance to where the tortoise had been, the tortoise will have moved
to a new location, ahead of Achilles. Will Achilles ever catch the tortoise?
What bothered Zeno and his friends was that it would require an infinite number of
mathematical steps of the type initially described. Could Achilles ever complete an infinite
number of such steps?
The problem, or paradox, posed by Zeno deals with limiting values, taking smaller and
smaller intervals. We can solve Zenos problem with algebra by requiring that Achilles and the
tortoise be at the same point at some time, t. The position of Achilles, at any time t, is
xA = x0 A + vA t
and the position of the tortoise is
xT = x0 T + vT t
and the initial positions differ by
x0T = x0A + 100
and because
vA = 10 vT
we find, by setting x0T = x0A, with this substitution for vA ,
9 vT t = 100 m
from which we can find when and where Achilles will catch the tortoise if we know v T (or v A).
Because each step is 1/10 as great as the previous step, the steps become infinitesimal and require
shorter and shorter times. Newton and others recognized this did not represent a real difficulty in
finding the sum of steps.
A quite different sort of question is the value of (x 2 - 4)/(x - 2) at the point x = 2. When x =
2
2, x - 4 = 0 and x - 2 = 0, so the ratio is 0/0, which is an indeterminate form. However, the
function is well-behaved; it approaches the same value, from above or from below.
6

2
0 1.75 3.5

Figure A1. Although the function (x2 - 4)/(x - 2) is indeterminate at


x = 2, it approaches the value x + 2 = 4 as x approaches 2 from
below or from above.

7/10/07 A- 120
In the following discussion, we can assume that the limits discussed are all well behaved. In
most instances, no difficulty arises at the specific values of interest. Even if a few of them are not
readily evaluated at a particular point, provided they appear well behaved on approaching that
point from below and from above, we will be justified in evaluating the limits by standard
methods.

A.3 Differential Calculus


An equation usually relates two or more variables, showing the values assumed by one
quantity as the other variable, or variables, take on different possible values. For example, the
pressure, volume, and temperature of an ideal gas are related by the equation

PV = nRT (A1)

in which n is the number of moles of gas, R is a universal constant (8.3144 J/molK, independent
of which real gas is being considered, to the approximation that the real gas follows this
equation), and the temperature is an absolute temperature, usually on the Kelvin scale.

Derivatives. One of the important questions that can be answered from such an equation
concerns the rate at which one variable changes with changes in another. For example, we may
ask how the volume changes with changes in temperature, for a fixed pressure. We can write

nR (A2a)
V1 = T1
P
(A2b)
nR
V2 = T2
P

and therefore
(A3)
nR nR
V = V 2 V1 = (T2 T1 ) = T
P P

or
V nR
= (A4)
T P

It can be seen from Figure A2 that this ratio is the slope of the line of volume plotted against
temperature.
Now suppose we are interested, instead, in how volume changes with pressure, at a fixed
temperature. The curve is a hyperbola, shown in Figure A3. Clearly the slope is no longer
constant. Proceeding as before, we may write

7/10/07 A- 121
V1 = nRT (1/P1) (A5a)
V2 = nRT (1/P2) (A5b)

Figure A2. Volume against


temperature for an ideal
gas. The slope of the line
is V/T = nR/P.

[Vertical axis V;
horizontal axis T.]

and subtracting, we obtain

1 1 nRT (P1 P2 ) -P
V = V 2 V1 = nRT = = nRT (A6)
P2 P1 P1 P2 P1 P2
or
V nRT
= (A7)
P P1 P2
This calculated value is not the slope of the curve at either P 1,V 1, or P 2 ,V 2 ; it is the slope of the
chord connecting these two points (b-d, Figure A3). Thus the slope depends not only on where
we start (P1,V1) but also on how far we go. If we want the slope at the point P1 ,V1 that is, the
slope of the line tangent to (touching) the curve at this point we can take P 2 closer and closer
to P1. If P2 is sufficiently close to P1 , we can write the
equation in the form
V nRT
= 2
P P = P1 = P2 P

(A8)

This says that the slope of the line tangent to the curve at
P1,V1 depends on P1, as it should by inspection of the
curve, but not on any other pressure value, which is also
quite reasonable.

7/10/07 A- 122
Figure A3. Volume against pressure for an ideal gas.
The slope of the chord, bd, is V/P = - nRT/P1P2.
The slope of the tangent at b, ac, is dV/dP = - nRT/P 12.

The slope of the line tangent to a curve is called the derivative of the curve, and is written in
the form dy/dx (or in this example, dV/dP). When we need to be more explicit (which is not very
often) we write

dy y (A9)
= Lim
dx x0 x
That is, dy/dx is the ratio of y/x as x becomes vanishingly small.
The two derivatives we have already met would thus be written

dV nR
= (A10)
dT P
and
dV nRT
= 2 (A11)
dP P

Derivatives cannot always be found as easily as for the two examples considered above, but for
present purposes only a very few formulas are required, and these few are given in Table A1.
From these few basic expressions it is possible to obtain many others. A few of these are listed in
Table A2. You should check each of these yourself, by applying the formulas from Table A1, to
be sure you see how the process works.
One word of warning. If you want the derivative at a particular point or for specific values
of the variables, do not substitute values of the variables first. Find the derivative, in terms of the
symbols, then substitute numbers for the symbols, as required.

Table A1 Basic Derivatives


Function Derivative
y=u+v dy/dx = du/dx + dv/dx

y = xn dy/dx = nxn-1

y = vn dy/dx = nvn-1 dv/dx

y = uv dy/dx = v du/dx + u dv/dx

7/10/07 A- 123
y = ln x dy/dx = 1/x

y = ev dy/dx = ev dv/dx

y = a (constant) dy/dx = 0

Although derivatives are customarily represented by notations such as dx/dt, proposed by


Leibniz, the short-hand notation of Newton is often preferred. Time derivatives are indicated by
the dot above; derivatives with respect to position by a prime. Thus

dy dy (A12)
y& = y' =
dt dx

Second derivatives are represented by double dots or primes.

d y& d2y dy ' d 2 y


&y& = = ; y" = = (A13)
dt dt 2 dx dx 2

Table A2 Additional Derivatives


Function Derivative
y = ax dy/dx = a

y = ax2 dy/dx = 2ax

y = a/x dy/dx = - a/x2


-(n+1)
y = ax-n dy/dx = -nax

y = u/v dy/dx =(1/v)du/dx - (u/v2) dvdx


= (1/v2)[ v du/dx - u dv/dx]
ax ax
y=e dy/dx = a e

y = x ln x dy/dx = ln x + 1

V = nRT/P dV/dT = nR/P (P constant)

E = mv2 dE/dv = mv

V = nRT/P dV/dP = -nRT/P2 (T constant)

7/10/07 A- 124
f = q1q2 /r2 df/dr = - 2q1q2 /r3

v f = vo2 + 2 ax (
dy / dx = 1 / 2 vo2 + 2 ax ) (2a )
1 / 2

= a / vo2 + 2 ax

Differentials. Although the derivatives are defined as the limiting value of a ratio, as the
bottom, and therefore the top, approach zero, it is also possible to interpret derivatives as a ratio
of two infinitesimal quantities -- that is, two quantities each of which is smaller than any number
you may select beforehand. When interpreted in this way, dy and dx are called differentials.

In the notation of differentials, we may write equations such as

(A14)
(P constant ) dV =
nR
dT
P

(T constant ) dV =
- nRT
dP (A15)
P2

dA = 2 r dr (A16)

dy (A17)
dy = dx
dx

A.4 Limits and Logarithms


Division is a basic operation, learned in elementary school. The division of a by b, a/b, is
equivalent to asking how many times we can subtract b from a (or, equivalently, by what quantity
must we multiply b to get a; b x ? = a). Thus 0/2 is well defined the answer is zero. But 2/0 is
undefined. You could remove zero from 2 all day long, and still have 2. Nor is there any definite
number c such that 0 x c = 2. We describe 2/0 as infinite, meaning that the answer is larger than
any number you might select beforehand.

Theory of Limits. Less easily analyzed is a division that is equivalent to 0/0. Is the answer 0?
Is the answer infinite, 4 ? Or can we obtain some meaningful value between zero and infinity?
You may be aware of proofs such as 2 = 1 that rely on a hidden division by zero. An important
segment of mathematics deals with the analysis of quantities that appear to be of the form of 0/0,
but can, on closer inspection, be assigned meaningful values.
Consider a similar problem of the theory of limits. As x becomes small, 1 + x approaches 1,
and 1 raised to any power (i.e., 1 multiplied by itself any number of times) would still give 1. But
1/x, as x becomes small, becomes very large, and any number greater than 1 raised to a

7/10/07 A- 125
sufficiently high power should give a large answer. What happens, then, to

(1 + x )1/ x
as x approaches zero? As you can show for yourself by substituting small values of x, the limiting
value of the expression is approximately 2.718, which we label as e.
Lim (1 + x)1/ x = e x60
x60

This quantity appears frequently, and quite naturally, in mathematical and physical problems. In
particular, it often appears as the base of an exponential expression or, equivalently, as a base of
logarithms. The number is irrational (like the familiar ), so it cannot be represented exactly in
decimal form; to 10 places it is e = 2.7182818285 ...

Logarithms. A logarithm is another name for an exponent. For example, we know that 10 3
= 1000. Therefore the logarithm of 1000, to base 10, is 3. The logarithm of 100, to base 10, is 2.
Choosing e as the base of logarithms, if

ez = w
then
lne w = ln w = z
ln uw = ln u + ln w (A18)
so
ln wn = n ln w
and
ln e = 1
where we choose the usual physicists notation

log10 x = log x loge x = ln x

From our definitions, it follows that, because

Lim (1 + x)1/x = e (A19)


x6 0

ln e = ln (1 + x)1/ x = 1/x ln (1 + x) = 1 x6 0
so

ln (1 + x) = x for small x. (A20)

This is very often a convenient approximation.

A.5 Summation by Integration


7/10/07 A- 126
Often we need to add together a large number of small changes in a variable or in some
expression involving variables. The summation process may be approached from either of two
viewpoints. One method, which we consider first, is represented geometrically by an area. The
second method may be called an antiderivative.

Area. In section A.3 we looked at the equation for an ideal gas to find how one quantity
changes as another quantity changes. That gave us the derivative, which may also be rewritten as
a ratio of differentials. Differentials may be equated, telling us more directly the infinitesimal
change in one quantity as some other quantity undergoes an infinitesimal change. A derivative, or
ratio of differentials, may always be interpreted as a slope.
For example, we found the derivative, dV/dT, when PV = nRT (and P is constant), to be

dV nR
= (A21)
dT P

and therefore
nR
( P constant) dV = dT (A22)
P

The right-hand side is a product of nR/P and the infinitesimal quantity, dT. Such a product may
be represented as in Figure A4. The ordinate is nR/P and dT is an infinitesimal change in the
abscissa. The product is an area the area of a vertical strip of infinitesimal width, as roughly
represented in the figure.

Figure A4. The product nR/P Tj is


the area of the jth rectangle. The sum
of all these rectangles (vertical slabs)
is the total volume change for T =
T1 to T2.

As the temperature changes, the location of the vertical strip changes, moving from left to
right for an increase in T. The quantity we seek is the sum of all these changes,
N
nR
P T
i
i (Ti 0, N )

7/10/07 A- 127
which is simply the area of the rectangle, between T 1 and T 2 . Because nR/P is constant (we
assumed constant pressure), the area is easily found. It is

nR
(T 2 T1 )
P

Similarly, by going through an equivalent process, we would find that

N
V = dV = Vi = V2 V1 (Vi 0, N ) (A23)
i

Therefore
V = V 2 V1 =
nR
(T 2 T1 ) = nR T (A24)
P P

Perhaps the analysis appears fussier than necessary, but the same method may now be
applied to a less obvious problem. We found if we hold temperature constant,

Figure A5. The product (-nRT/Pj2 )Pj


is the area of the jth rectangle. The
sum of these approximates the area
under the curve between P1 and P2.

dV nRT
= (A25)
dP P2
or
dP
dV = nRT (A26)
P2
As shown in Figure A5, the summation procedure gives (for T constant),

7/10/07 A- 128
N N
Pi
V = Vi = nRT ( Vi , Pi 0; N ) (A27)
i i Pi 2

We can evaluate the right-hand side easily enough with a computer or a programmable hand
calculator. For the initial value, P = P1, calculate V1 = nRT/P1. Pick some arbitrary, small value
for P. Then for P = P1 + P, calculate a new V = V, and from this V = V - V1 . Continue
adding P to find P = P + P, then V and hence V = V - V, and so forth. The result, as
depicted in Figure A5, gives V as the area under the curve of nRT/P 2 vs. P.
The accuracy of the answer will depend, in general, on the size of the P we choose. A
smaller value of P will give a more accurate result, but may take a little more computer time. In
principle, we could also find the area, and thus V, by cutting out the shaded area and weighing
it, comparing the weight to a rectangle of the same paper of known area.

Antiderivative. It is often (but not always) possible to find a formula to express the answer
to summation problems such as those just considered. A clue is offered by the last sum, which
can be written as a sum of infinitesimal terms of the form

nRT
d (V ) = d
dP
= nRT 2 (A28)
P P

The sum of small changes is the total change. Writing the sum (G, Greek S) with a stylized S, I,
to represent the sum over an infinite number of infinitesimal quantities, we can express the
problem in the form

dP nRT
dV = nRT P2
= d P
(A29)

and therefore, just as summing up all the small changes in V gives V, summing up all the small
changes in nRT/P gives (nRT/P), so

nRT
V = (A30)
P

where we have taken advantage of the knowledge (Table A2) that

dP nRT
nRT = d (A31)
P2 P

7/10/07 A- 129
That is, we recognize where - nRT/P2 came from, which is precisely the reverse problem of
finding the differential (or the derivative) in the first place. Therefore this process has been called
taking the antiderivative.
We call IdV and I-nRT dP/P2 and I (nR/P) dT integrals. The expression following the
integration sign, I, is called the integrand. The approximation errors of numerical integration
depend on the size of the steps, and can therefore generally be made negligible. An analytical
solution fits the curve at each point of the curve, so there is no error of approximation.
To evaluate an integral we ask, What function could we take the differential of to obtain
this integrand? Some basic integrals we typically learn to remember. Extensive tables are
available to help us with more complex integrations. A few examples are given in Table A3.

Table A3. Basic Integrals

Definite and Indefinite Integrals. Every physical problem is subject to boundary conditions,
meaning in this instance that we find the sum, or integral, or total change, between a definite
initial state and a definite final state. We express this by noting the limiting values in writing the
integral and in evaluating the result. Thus we would write

nRT dP nRT 2
P
V2 P2
V1
dV =
P1 P2
=
P P1 (A32)

which gives
1 1
V 2 V1 = nRT (A33)
P
2 P1

These are called definite integrals.

7/10/07 A- 130
There are times, however, when we are more interested in finding the function from which
the integrand came (i.e., finding the antiderivative) than in finding the actual change. Then we
must recognize that the derivative of a constant is zero, so any arbitrary constant added to the
integral vanishes in finding the integrand. Thus we may write

dP nRT
nRT P 2
=
P
+c (A34)

where c is an arbitrary, or unknown, constant. Such an expression is called an indefinite integral.

7/10/07 A- 131

You might also like