Thermodynamics Notes
Thermodynamics Notes
The solutions which obey Raoult’s law over the entire range of concentration are known as
ideal solutions. The ideal solutions have two other important properties. The enthalpy of
mixing of the pure components to form the solution is zero and the volume of mixing is
also zero, i.e.,
∆mixH = 0, ∆mixV = 0
It means that no heat is absorbed or evolved when the components are mixed. Also, the
volume of solution would be equal to the sum of volumes of the two components. At
molecular level, ideal behavior of the solutions can be explained by considering two
components A and B. In pure components, the intermolecular attractive interactions will be
of types A-A and B-B, whereas in the binary solutions in addition to these two interactions,
A-B type of interactions will also be present. If the intermolecular attractive forces between
the A-A and B-B are nearly equal to those between A-B, this leads to the formation of ideal
solution. A perfectly ideal solution is rare but some solutions are nearly ideal in behavior.
Solution of n-hexane and n-heptane, bromoethane and chloroethane, benzene and toluene,
etc. fall into this category.
When a solution does not obey Raoult’s law over the entire range of concentration, then it is
called non-ideal solution. The vapour pressure of such a solution is either higher or lower
than that predicted by Raoult’s law. If it is higher, the solution exhibits positive deviation and
if it is lower, it exhibits negative deviation from Raoult’s law.
Let us consider a binary solution of two volatile liquids and denote the two components as 1
and 2. When taken in a closed vessel, both the components would evaporate and eventually
an equilibrium would be established between vapour phase and the liquid phase. Let the total
vapour pressure at this stage be P total and P1 and P2 be the partial vapour pressures of the two
components 1 and 2 respectively. These partial pressures are related to the mole fractions x1
and x2 of the two components 1 and 2 respectively.
The French chemist, Francois Marte Raoult (1886) gave the quantitative relationship between
them. The relationship is known as the Raoult’s law which states that for a solution of
2
volatile liquids, the partial vapour pressure of each component of the solution is directly
proportional to its mole fraction present in solution.
In chemical process plants – the ultimate domain of application of the principles of chemical
engineering thermodynamics – matter is dominantly processed in the form of mixtures.
Process streams are typically comprised of multiple components, very often distributed over
multiple phases. Separation or mixing processes necessitate the use of multiple phases in
order to preferentially concentrate the desired materials in one of the phases. Reactors very
often bring together various reactants that exist in different phases. It follows that during
mixing, separation, inter-phase transfer, and reaction processes occurring in chemical plants
multi-component gases or liquids undergo composition changes. Thus, in the thermodynamic
description of such systems, in addition to pressure and temperature, composition plays a key
role.
Further, whenever multiple phases are present in a system, material and energy transfer
occurs between till the phases are in equilibrium with each other, i.e., the system tends to a
state wherein the all thermal, mechanical and chemical potential gradients within and across
all phases cease to exist.
The present chapter constitutes a systematic development of the concept of a new class of
properties essential to description of real mixtures, as well of the idea of the chemical
potential necessary for deriving the criterion of phase and chemical reaction equilibrium.
Such properties facilitate the application of the first and second law principles to
quantitatively describe changes of internal, energy, enthalpy and entropy of multi-component
and multiphase systems. Of the separate class of properties relevant to multi-component and
multi-phase systems, the partial molar property and the chemical potential are particularly
important.
The former is used for describing behavior of homogeneous multi-component systems, while
the latter forms the fundament to description equilibrium in multi-phase, as well as reactive
3
systems. As in the case of pure gases, the ideal gas mixture acts as a datum for estimating the
properties of real gas mixtures. The comparison of the properties of the real and ideal gas
mixtures leads to the introduction of the concept of fugacity, a property that is further related
to the chemical potential. Fugacity may also be expressed as a function of volumetric
properties of fluids. As we will see, the functional equivalence of fugacity and the chemical
potential provides a convenient pathway for relating the temperature, pressure and phase
composition of a system under equilibrium.
X X X
dX = dn1 + dn2 + dn3 + ...
n1 n2 n3
This just says that any change in X will be given by how much the amounts of components 1,
2, 3… are changed (dni) and by the coefficient that tells us what the amount of X is per mole
X
of substance 1, 2, 3…, . These coefficients are called partial molar quantities. For
ni
small changes we can often integrate this expression and obtain:
X X X
X = n1 + n2 + n3 + ...
n1 n2 n3
Lets consider an example. One common partial molar quantity is the partial molar volume.
V
The partial molar volume of component i is just given by Vi = . Thus it is the
ni T , P ,n'
amount that the volume would change if we changed the number of moles of substance i on a
per mole basis. The simplest case is a mixture of ideal gases, call them A and B. Note that in
this case we hold T and P constant as well as the amounts of all other substances in the
4
mixture (n'). We know that ideal gases do not interact and that their volumes depend only on
n A RT nB RT V RT
how much is present at constant T and P: V = + . Thus = . So it
P P n A T , P ,n ' P
turns out that the partial molar volume of any component in a mixture of ideal gases is RT/P.
It is not always that simple, but more on that later.
OK, now we come to the really important partial molar quantity -- the partial molar Gibbs
free energy. Just as for volume, we can write:
G G
G = n1 + n2 + ...
n1 T , P ,n ' n2 T , P ,n '
G G
1 = , 2 = ,...
1 T , P ,n '
n
2 T , P ,n '
n
Thus
G = 1 n1 + 2 n 2 + ...
and
dG = 1 dn1 + 2 dn2 + ...
At constant temperature and pressure and under conditions where the chemical potentials are
constant with changing composition. In fact, we can just add these new terms to our previous
expression for dG at constant pressure and constant n:
which reduces to the equation just above this if the temperature and pressure are constant.
An aside…
G = 1 n1 + 2 n2 + ...
and
dG = 1 dn1 + 2 dn2 + ... + d1 n1 + d 2 n2 + ...
5
Yet from the arguments made above, we can see that this equation must be the same as the
previous equation for dG:
Therefore
0 = +d1n1 + d 2 n2 + ...
This is an important equation called the Gibbs-Duhem equation. What it says is that any
change in the chemical potential of one component of a closed system must be balanced by
another.
Applications
(i) Gibbs-duhem equation is helpful in calculating partial molar quantity of a binary mixture
by measuring the composition of the mixture which depends on the total molar quantity.
(ii) Gibbs-duhem equation is helpful in calculating the partial vapor pressures by calculating
the total vapor pressure. All these calculations require a curve-fitting procedure. Using
tabulated experimental data the accuracy of the calculated quantities was found to be
comparable to the accuracy of the original experimental data.
• If two systems have the same temperature there is no net energy flow between them.
• If two systems have the same pressure, there is no net change of volumes
• If two systems have the same X, there is no net flow of particles What is this X? We call it
the chemical potential:
• to redistribute in space,
can be expressed by one and the same quantity – its chemical potential µ.
By definition the partial molar Gibbs free energy is termed the chemical potential of species i
in the mixture, i.e.,:
6
1.5.1 Effect of temperature and pressure on chemical potential
The above relation is known as the Lewis/Randall rule, and applies to each species in an
ideal solution at all conditions of temperature, pressure, and composition. It shows that
the fugacity of each species in an ideal solution is proportional to its mole fraction; the
7
proportionality constant being the fugacity of pure species i in the same physical state as
the solution and at the same T and P.
Henry was the first to give a quantitative relation between pressure and solubility of a gas in a
solvent which is known as Henry’s law. The law states that at a constant temperature, the
solubility of a gas in a liquid is directly proportional to the partial pressure of the gas present
above the surface of liquid or solution.
p = KH x
p
= o + RT ln , (1)
po
For a mixture of ideal gases it can be shown that for each component, i, the chemical
potential is given by,
pi
i = io + RT ln . (2)
po
For a nonideal gas we used the fugacity, f, instead of the pressure and the chemical potential
for one component is,
f ( p)
= o + RT ln . (3)
po
fi
i = io + RT ln , (4)
po
8
only now the fugacity of component i is a function of the pressures of all the gases in the
mixture,
fi = f i ( p1 , p2 , p3 ,K ). (5)
i = io + RT ln X i . (6)
These expressions for chemical potential all have the form of a reference or standard state
chemical potential plus RT times the logarithm of something related to pressure or
concentration. This form turns out to be very important, so important that G. N. Lewis used
it to give the most general case of chemical potential as
i = io + RT ln ai . (7)
The quantity, ai, is called the "activity" of component i and Equation 7 should be regarded as
the definition of activity. Notice that the activity has no units.
All of the special cases we have been considering so far can be reconciled to this definition of
activity. Thus, for an ideal gas mixture,
pi
ai = ;
po
fi
ai = ;
po
ai = X i ,
and so on. In a nonideal solution we would have to just write Equation 7 again,
i = io + RT ln ai . (7)
9
In Equation 7 all the non idealities of the solution are absorbed into the activity. We will see
a more convenient way to write this below under the heading "activity coefficient."
We can find the activity of a component of a nonideal solution from measurements of the
vapor pressure of that component in the vapor in equilibrium with the solution. We know that
the chemical potential of a component must be the same in the vapor as in the liquid. that is,
from Equations 2 and 7 we obtain,
il = ig
pi (8a, b)
il* + RT ln ai = igo + RT ln ,
po
pi*
il* = igo + RT ln . (9)
po
pi pi*
RT ln ai = RT ln o − RT ln o
p p
(10a, b)
p
= RT ln *i ,
pi
so that
pi
ai = . (11)
pi*
If the solution were ideal pi would be given by Raoult's law and the activity would be just the
mole fraction.
Activity Coefficient
Sometimes it is convenient to write the activity as the product of an ideal part times a non
ideality correction part. For example, in a real solution we might write,
ai = X i i , (12)
10
f i = pi i , (13)
so that
fi pi i pi
ai = = = o i. (14)
po po p
p
( Z −1)
dp '
f ( p) = pe
p'
0
, (15)
so that
= e
p
( Z −1)
dp '
p'
0
. (16)
pi
ai = = X i i , (17)
pi*
In general, the activity coefficient is a unit less parameter that contains all of the nonideality
of a system.
Remember for ideal gases that the chemical potential was given by:
11
PA
A ( g ) = A + RT ln
P
Gmix = nRT ( X A ln X A + X B ln X B )
where the mole fractions are mole fractions in the gas phase.
A (l ) = *A (l ) + RT ln( X A )
and it should not surprise you to learn that the Gibbs free energy of mixing is just
Gmix = nRT ( X A ln X A + X B ln X B )
In other words, the Gibbs free energy of mixing two ideal gases is the same as the Gibbs free
energy for forming an ideal solution of two liquids. For more components, you just add more
terms. The entropy and enthalpy of mixing are also the same as with ideal gases.
S mix = −nR( X A ln X A + X B ln X B )
Remember, these equations only apply to ideal solutions! In real situations, the enthalpy, in
particular, is usually not zero.
Let's just see what effects a mildly nonideal case would have on these (in the book, this is
partly covered at the end of the chapter under activities). We can see that for ideal dilute
solutions (Henry's law) instead of just XA in the logarithm, we would have something
proportional to XA. As we saw before:
X AK A
A (l ) = *A (l ) + RT ln *
PA
12
Gi = n A A + n B B
= n A A + nB B
* *
Now for a ideal dilute mixture where there is just a wee bit of A mixed in with a lot of B:
G f = n A A + nB B
X K
= n A *A (l ) + RT ln A * A + n B B* + RT ln( X B )
PA
X K
= n A A + n A RT ln A * A + n B B + n B RT ln X B
* *
PA
KA
= n A A + n A RT ln + n A RT ln X A + n B B + n B RT ln X B
* *
*
PA
Now subtract the initial from the final Gibbs free energy:
KA
Gmix = G f − Gi = n A A + n A RT ln + n A RT ln X A + n B B + n B RT ln X B − n A A − n B B
* * * *
*
PA
KA
Gmix = n A RT ln + n A RT ln X A + n B RT ln X B
PA*
K
Gmix = nRT X A ln *A + X A ln X A + X B ln X B
PA
or
KA
Gmix = nRTX A ln + nRT ( X A ln X A + X B ln X B )
PA*
How this extra term partitions itself between enthalpy and entropy depends on the details of
the liquid interactions and structure. However, usually the bulk of the extra term ends up
making the enthalpy of mixing nonzero.
• McCabe Thiele method assumes constant molar flow rate because it considers equal
latent heat of vaporization.
• Here we consider varying molar flow rate by solving simultaneous material and
energy balances.
• In this case, the operating lines for the enriching and stripping section will be
determined from simultaneous solution of mass and energy balance equations.
13
• To facilitate the solution of the heat balance equation, an enthalpy diagram can be
constructed and used.
14
1.13 Excess Properties
Unlike for real gases (pure or mixtures) the EOS based approach to calculation of
thermodynamic properties of real liquid solutions have not proved very successful. However,
as molar residual property is defined for real gases, for real liquid solutions one may
formulate a different departure function called the molar excess property that quantify the
deviation from ideal solution property. The mathematical formalism of excess properties is,
therefore, analogous to that of the residual properties. If M represents the molar (or unit-
mass) value of any extensive thermodynamic property(e.g., V,U, H, S, G, etc.), then an excess
property is defined as the difference between the actual property value of a solution and the
value it would have as an ideal solution at the same temperature, pressure, and composition.
Thus: M
The excess property bear a relationship to the property change of mixing. One may take the
example of excess Gibbs free energy to illustrate the point
The non-ideality of real liquid solutions are depicted well by use of excess properties,
especially through the behavior of The excess Gibbs energy is typically obtained from low
pressure vapour-liquid equilibrium data, while HE is obtained by measuring isothermal
enthalpy change of mixing. Lastly SE is derived using the following relation:
it shows the variation of each of the excess property as a function of liquid mole fraction for
a number of binary solutions.
15
1.14 Relation between excess Gibbs free energy and activity coefficient
This choice is in principle arbitrary, but the following standard states are usually chosen for
practical reasons:
• Mixtures of gases For components in the gaseous phase, the standard state of an ideal gas at
standard pressure is used. In the case of this choice, µist is the chemical potential of a pure
substance i in the state of an ideal gas at the temperature of the mixture and standard pressure,
usually pst = 101 325 Pa.
• Mixtures of liquids and solid substances
For components in a liquid or solid phase, the standard state of a pure substance at the
temperature and pressure of the system is used. µist is the chemical potential of a pure
substance i, which at the temperature and pressure of the mixture is in the same phase as the
mixture.
16
The standard states of individual components may be chosen arbitrarily within certain limits.
The first limitation is that the standard temperature is always the temperature of the system.
Further on, the standard state of individual components of a system is usually chosen at fixed
composition. The standard pressure is either constant, or it changes with the system pressure.
• Standard state: the state of a pure substance in a given state of matter or modification at the
temperature of the system and a standard pressure.
Solved problems
1.1
1.2
17
1.3
REFERENCES
18
SCHOOL OF BIO AND CHEMICAL
DEPARTMENT OF CHEMICAL ENGINEERING
The last two criteria are better understood in terms of the property called chemical potential,
which we shall introduce in the following section. Here we focus on the overall, general criterion
that must be in obeyed for any chemical system in equilibrium regardless of whatever species,
phases or reactions that defines it. Consider a closed system, which can be either homogeneous
or heterogeneous, and which exists in a state of thermal and mechanical equilibrium with its
surroundings. However, we assume that it is not under equilibrium with respect to possible inter-
phase transfer of the chemical species or reactions between them. If the latter conditions prevail,
all inter-phase transfer processes or reactive transformation of species must continue to occur till
the point when the system is also at chemical equilibrium. In all real systems such changes are
induced by finite gradients and are therefore are irreversible in nature. Applying the first law
equation for all such changes for the system.
We next consider that the system is under thermal and mechanical equilibrium with the
surroundings. Under such a situation,
2
Combining the above equations,
It follows that for all incremental changes within the system, which take it closer to the final
thermodynamic equilibrium, the property changes must satisfy the constraint imposed by the
above eqn. If the changes internal to the system occur under reversible conditions, the equality
sign is valid; on the other hand, for irreversible processes the inequality condition holds. The
above equation can be used to generate alternate criteria of thermodynamic equilibrium,
namely:
The other two criteria which are most apt in relation to thermodynamics of phase equilibria
involve the use of Helmholtz and Gibbs free energy. If a process takes place under the
constraints of constant temperature and volume then:
And, if the process occurs under constant temperature and pressure one may write:
This equation provides the most practical of the three versions of general criteria of approach to
equilibrium, as temperature and pressure are the most easily measurable of all the
thermodynamic properties. As we have argued at the early part of this section at total
thermodynamic equilibrium not only are thermal and mechanical gradients non-existent, there
3
can be no further change in either the composition of any of the phases, or that of the reactive
species. If there are any such incremental, infinitesimal changes of composition variables at the
state of complete equilibrium the system once again must return to its stable state. the following
equation to characterize thermodynamic equilibrium:
4
two phases. To generalize the results we assume that any two types of phases a and b of a pure
component are at equilibrium. Thus as given by equation
However, for a pure component the chemical potential is reduces to the pure component molar
Gibbs free energy. Therefore: and,
On taking a differential:
Using the generic relationship ,we may write in keeping with the fact that for a given equilibrium
temperature, the equilibrium pressure corresponds to the saturation vapour pressure :
On rearranging:
Thus, ,
The last equation is called the Clapeyron equation. For the specific case of phase transition from
liquid (l) to vapor (v), it translates into:
5
Noting that liquid phase molar volumes are relatively much lesser than vapour phase volumes,
we may write,
Further at low to moderate saturation pressures if we assume ideal vapor phase behavior, then
Or:
6
2.2.1 Fugacity of Pure substances:
It has been shown that the chemical potential provides a fundamental description of phase
equilibria. As we shall further, it also proves an effective tool for depicting chemical reaction
equilibria. Nevertheless, its direct usage is restricted, as it is not easy to directly relate the
chemical potential to thermodynamic properties amenable to easy experimental determination,
such as the volumetric properties. The definition of a new function called fugacity, itself related
to the chemical potential, helps bridge the gap. The concept of fugacity is advanced based on the
following thermodynamic relation for an ideal gas. For a single component closed system
containing an ideal gas we have:
At constant temperature, for a pure ideal gas ‘i’ the above equation reduces to:
(At const T)
Thus, ;
Hence
Since fi has the units of pressure, it is often described as a “fictitious pressure”. It may be noted
that the definition of fugacity as provided and is completely general in nature, and so can be
extended to liquids and solids as well.
7
Thus:
Where,
However, by equation :
Thus, using the last relation
(At const. T)
The general criterion of thermodynamic equilibrium has been defined by eqn. Applying it to, for
example, a vapour (V) and liquid (L) system of a pure component ‘i’ we have:
Thus: and
The above equation may be generalized for any other types of phases. However, the eqn. is
rendered more easily applicable if the chemical potential is replaced by fugacity. Thus
integrating eqn. between vapour and liquid states of a pure component:
In eqn. indicates the value for either saturated liquid or saturated vapor, this is because the
coexisting phases of saturated liquid and saturated vapor are in equilibrium. Since under such
8
condition the pressure is we can write:
Fugacity coefficient (and hence fugacity) of pure gases may be conveniently evaluated by
applying eqn. to a volume-explicit equation of state. The truncated virial EOS is an example of
the latter type, for which the compressibility factor of pure species (i) is given by:
Or
9
10
Vapor pressure and boiling
The vapor pressure of a liquid at a particular temperature is the equilibrium pressure exerted by
molecules leaving and entering the liquid surface.
11
• Distillation occurs because of the differences in the volatility of the components in the
liquid mixture.
THE BOILING POINT DIAGRAM
• The Boiling Point Diagram The boiling point diagram shows how the equilibrium
compositions of the components in a liquid mixture vary with temperature at a fixed
pressure.
• Consider an example of a liquid mixture containing 2 components (A and B) - a binary
mixture..
The boiling point of A is that at which the mole fraction of A is 1. The boiling point of B
is that at which the mole fraction of A is 0. In this example, A is the more volatile
component and therefore has a lower boiling point than B.
The upper curve in the diagram is called the dew-point curve while the lower one is
called the bubble-point curve.
12
• For example, when a subcooled liquid with mole fraction of A=0.4 (point A) is heated, its
concentration remains constant until it reaches the bubble-point (point B), when it starts
to boil. The vapors evolved during the boiling has the equilibrium composition given by
point C, approximately 0.8 mole fraction A. This is approximately 50% richer in A than
the original liquid. ™
• This difference between liquid and vapor compositions is the basis for distillation
operations.
Relative Volatility
αij =
Distillation columns are designed based on the boiling point properties of the components in the
mixtures being separated. Thus the sizes, particularly the height, of distillation columns are
determined by the vapor liquid equilibrium (VLE) data for the mixtures.
Constant pressure VLE data is obtained from boiling point diagrams. VLE data of binary
mixtures is often presented as a plot, as shown in the figure. The VLE plot expresses the bubble-
point and the dew-point of a binary mixture at constant pressure. The curved line is called the
13
equilibrium line and describes the compositions of the liquid and vapour in equilibrium at some
fixed pressure
The previous particular VLE plot shows a binary mixture that has a uniform vapor-liquid
equilibrium that is relatively easy to separate. The next two VLE plots below on the other hand,
shows non-ideal systems which will present more difficult separations.
14
The most intriguing VLE curves are generated by azeotropic systems. An azeotrope is a liquid
mixture which when vaporised, produces the same composition as the liquid. The two VLE plots
below, show two different azeotropic systems, one with a minimum boiling point and one with a
maximum boiling point. In both plots, the equilibrium curves cross the diagonal lines, and this
are azeotropic points where the azeotropes occur. In other words azeotropic systems give rise to
VLE plots where the equilibrium curves crosses the diagonals
Raoult's law
It states that the partial vapor pressure of each component of an ideal mixture of liquids is equal
to the vapor pressure of the pure component multiplied by its mole fraction in the mixture.
where is the partial vapor pressure of the component in the gaseous mixture (above the
solution), is the vapor pressure of the pure component , and is the mole fraction of the
component in the mixture (in the solution).
15
Once the components in the solution have reached equilibrium, the total vapor pressure of the
solution can be determined by combining Raoult's law with Dalton's law of partial pressures to
give
.
If a non-volatile solute (zero vapor pressure, does not evaporate) is dissolved into a solvent to
form an ideal solution, the vapor pressure of the final solution will be lower than that of the pure
solvent.
Ideal solution: An ideal solution will obey Raoult's Law,
Real solutions: Solutions which deviate from Raoults law are called as real solutions. Many pairs
of liquids are present in which there is no uniformity of attractive forces, i.e.,
the adhesive and cohesive forces of attraction are not uniform between the two liquids, so that
they deviate from the Raoult's law.
16
Negative deviation:
If the vapor pressure of a mixture is lower than expected from Raoult's law, there is said to be
a negative deviation. This is evidence that the adhesive forces between different components are
stronger than the average cohesive forces between like components. In consequence each
component is retained in the liquid phase by attractive forces that are stronger than in the pure
liquid so that its partial vapor pressure is lower.
For example, the system of chloroform (CHCl3) and acetone(CH3COCH3) has a negative
deviation from Raoult's law
Positive deviation
When the cohesive forces between like molecules are greater than the adhesive forces between
dissimilar molecules, the dissimilarities of polarity leads both components to escape solution
more easily. Therefore, the vapor pressure is greater than expected from the Raoult's law,
showing positive deviation. If the deviation is large, then the vapor pressure curve shows a
maximum at a particular composition and form a positive azeotrope. Some mixtures in which
this happens are (1)benzene and methanol, (2) carbon disulfide and acetone, and
(3) chloroform and ethanol.
The concentration of a vapor in contact with its liquid, especially at equilibrium, is often
expressed in terms of vapor pressure, which will be a partial pressure (a part of the total gas
pressure) if any other gas(es) are present with the vapor. The equilibrium vapor pressure of a
liquid is in general strongly dependent on temperature. At vapor–liquid equilibrium, a liquid with
individual components in certain concentrations will have an equilibrium vapor in which the
concentrations or partial pressures of the vapor components have certain values depending on all
of the liquid component concentrations and the temperature.
The converse is also true: if a vapor with components at certain concentrations or partial
pressures is in vapor–liquid equilibrium with its liquid, then the component concentrations in the
liquid will be determined dependent on the vapor concentrations and on the temperature. The
equilibrium concentration of each component in the liquid phase is often different from its
concentration (or vapor pressure) in the vapor phase, but there is a relationship. The VLE
17
concentration data can be determined experimentally, or computed or approximated with the
help of theories such as Raoult's law, Dalton's law, and Henry's law.
Such VLE information is useful in designing columns for distillation, especially fractional
distillation, which is a particular specialty of chemical engineers. Distillation is a process used to
separate or partially separate components in a mixture by boiling (vaporization) followed
by condensation. Distillation takes advantage of differences in concentrations of components in
the liquid and vapor phases.
In mixtures containing two or more components, the concentrations of each component are often
expressed as mole fractions. The mole fraction of a given component of a mixture in a particular
phase (either the vapor or the liquid phase) is the number of moles of that component in
that phase divided by the total number of moles of all components in that phase. Binary mixtures
are those having two components. Three-component mixtures are called ternary mixtures. There
can be VLE data for mixtures with even more components, but such data is often hard to show
graphically. VLE data is a function of the total pressure, such as 1 atm or whatever pressure the
process is conducted at.
When a temperature is reached such that the sum of the equilibrium vapor pressures of the liquid
components becomes equal to the total pressure of the system (it is otherwise smaller), then
vapor bubbles generated from the liquid begin to displace the gas that was maintaining the
overall pressure, and the mixture is said to boil. This temperature is called the boiling point of
the liquid mixture at the given pressure. (It is assumed that the total pressure is held steady by
adjusting the total volume of the system to accommodate the specific volume changes that
accompany boiling.) The boiling point at an overall pressure of 1 atm is called the normal
boiling point of the liquid mixture. Binary mixture VLE data at a certain overall pressure, such as
1 atm, showing mole fraction vapor and liquid concentrations when boiling at various
temperatures can be shown as a two-dimensional graph called a boiling-point diagram. The
mole fraction of component 1 in the mixture can be represented by the symbol x1. The mole
fraction of component 2, represented by x2, is related to x1 in a binary mixture as follows:
18
Boiling-point diagram
The preceding equilibrium equations are typically applied for each phase (liquid or vapor)
individually, but the result can be plotted in a single diagram. In a binary boiling-point diagram,
temperature (T ) is graphed vs. x1. At any given temperature where both phases are present, vapor
with a certain mole fraction is in equilibrium with liquid with a certain mole fraction. The two
mole fractions often differ. These vapor and liquid mole fractions are represented by two points
on the same horizontal isotherm (constant T ) line. When an entire range of temperatures vs.
vapor and liquid mole fractions is graphed, two (usually curved) lines result. The lower one,
representing the mole fraction of the boiling liquid at various temperatures, is called the bubble
point curve. The upper one, representing the mole fraction of the vapor at various temperatures,
is called the dew point curve.
These two lines (or curves) necessarily meet where the mixture becomes purely one component,
namely where x1 = 0 (and x2 = 1, pure component 2) or x1 = 1 (and x2 = 0, pure component 1).
The temperatures at those two points correspond to the boiling points of each of the two pure
components.
For certain pairs of substances, the two curves also coincide at some point strictly between x1 =
0 and x1 = 1. When they meet, they meet tangently; the dew-point temperature always lies above
the boiling-point temperature for a given composition when they are not equal. The meeting
19
point is called an azeotrope for that particular pair of substances. It is characterized by an
azeotrope temperature and an azeotropic composition, often expressed as a mole fraction. There
can be maximum-boiling azeotropes, where the azeotrope temperature is at a maximum in the
boiling curves, or minimum-boiling azeotropes, where the azeotrope temperature is at a
minimum in the boiling curves.
If one wants to represent a VLE data for a three-component mixture as a boiling point "diagram",
a three-dimensional graph can be used. Two of the dimensions would be used to represent the
composition mole fractions, and the third dimension would be the temperature. Using two
dimensions, the composition can be represented as an equilateral triangle in which each corner
represents one of the pure components. The edges of the triangle represent a mixture of the two
components at each end of the edge. Any point inside the triangle represents the composition of a
mixture of all three components. The mole fraction of each component would correspond to
where a point lies along a line starting at that component's corner and perpendicular to the
opposite edge. The bubble point and dew point data would become curved surfaces inside a
triangular prism, which connect the three boiling points on the vertical temperature "axes". Each
face of this triangular prism would represent a two-dimensional boiling-point diagram for the
corresponding binary mixture. Due to their three-dimensional complexity, such boiling-point
diagrams are rarely seen. Alternatively, the three-dimensional curved surfaces can be represented
on a two-dimensional graph by the use of curved isotherm lines at graduated intervals, similar to
iso-altitude lines on a map. Two sets of such isotherm lines are needed on such a two-
dimensional graph: one set for the bubble point surface and another set for the dew point surface.
It implies that for any closed system formed initially from given masses of a number of chemical
species, the equilibrium state is completely determined when any two independent variables are
fixed. The two independent variables that one may choose to specify may be either intensive or
extensive. However, the number of independent intensive variables is given by the phase rule.
Therefore, it follows that when F = 1, at least one of the two variables must be extensive, and
when F = 0, both must be extensive
20
2.6 Pxy and Txy diagrams.
21
Solved problems
2.1
22
2.2
REFERENCES
1. Smith J.M. and Van Ness H.C., Introduction to Chemical Engineering Thermodynamics,
7th Edition, McGraw Hill, 2005.
2. Narayanan K.V., A Text Book of Chemical Engineering Thermodynamics, 3rd Edition,
Prentice Hall of India Pvt. Ltd.,
2013.
3. Gopinath Halder, Introduction to Chemical Engineering Thermodynamics, 2 nd Edition,
PHI Learning Pvt. Ltd., 2009.
23
SCHOOL OF BIO AND CHEMICAL
DEPARTMENT OF CHEMICAL ENGINEERING
✓ A condition where a liquid phase and vapor phase are in equilibrium with each other
✓ e.g: mixture of liquid and vapor at an equilibrium level takes place when liquid and
vapor are allowed to contact to each other in a closed location
2
✓ Upper surface- sat. L states (P-T-x1)
✓ Liquid at F, reduces pressure at constant T & composition along FG, the first bubble
appears at L – bubble point: a point when a liquid forms the first bubble of vapor and
begins to evaporate
✓ As pressure reduces, more & more L vaporizes until completed at W; point where last
drop of L (dew) disappear – dew point: a point when a vapor forms the first droplet of
liquid and begins to condense
Assumptions;
Valid only if the species are chemically similar (size, same chemical nature e.g.
isomers such as ortho-, meta- & para-xylene)
yi P = xi Pi sat (i = 1,2,..., N )
xi : L − phase mole fraction
yi : V − phase mole fraction
Pi sat : Vapor pressure of pure species i
P : Total pressure
4
3.4 NRTL
3.5 UNIQUAC
UNIQUAC (short for UNIversal QUAsi Chemical model) builds on the work of Wilson by
making three primary refinements. First, the temperature dependence of Wij is modified to
depend on surface areas rather than volumes, based on the hypothesis that the interaction
energies that determine local compositions are dependent on the relative surface areas of the
molecules. If the parameter qi is proportional to the surface area of molecule i,
3.6 UNIFAC
This is an extension of UNIQUAC with no adjustable parameters for the user to input or fit to
5
experimental data. Instead, all of the adjustable parameters have been characterized by the
developers of the model based on group contributions that correlate the data in a very large data
base. The assumptions regarding coordination numbers, etc., are similar to the assumptions in
UNIQUAC. The same strategy is applied,
The Van Laar equation is an activity model, which was developed by Johannes van Laar in
1910-1913, to describe phase equilibria of liquid mixtures. The equation was derived from the
Van der Waals equation. The original van der Waals parameters didn't give good description of
vapor-liquid phase equilibria, which forced the user to fit the parameters to experimental results.
Because of this, the model lost the connection to molecular properties, and therefore it has to be
regarded as an empirical model to correlate experimental results.
6
Dewpoint & Bubblepoint Calculations with
Raoult’s Law
y i i =1
P = xi Pi sat
i
(
P = P2sat + P1sat − P2sat x1)
x1 P1 sat
y1 =
P
7
Raoult’s law equation can be solved for xi
to solve for dewpoint calculation (T is
given)
i xi =1
P=
1 1
y sat P=
i
i Pi
y1 / P1 sat + y2 / P2sat
y1 P
x1 =
P1 sat
The vapor pressure of a liquid at a particular temperature is the equilibrium pressure exerted by
molecules leaving and entering the liquid surface.
8
Boiling-point diagram
The preceding equilibrium equations are typically applied for each phase (liquid or vapor)
individually, but the result can be plotted in a single diagram. In a binary boiling-point diagram,
temperature (T ) is graphed vs. x1. At any given temperature where both phases are present, vapor
with a certain mole fraction is in equilibrium with liquid with a certain mole fraction. The two
mole fractions often differ. These vapor and liquid mole fractions are represented by two points
on the same horizontal isotherm (constant T ) line. When an entire range of temperatures vs.
vapor and liquid mole fractions is graphed, two (usually curved) lines result. The lower one,
representing the mole fraction of the boiling liquid at various temperatures, is called the bubble
point curve. The upper one, representing the mole fraction of the vapor at various temperatures,
is called the dew point curve.
These two lines (or curves) necessarily meet where the mixture becomes purely one component,
namely where x1 = 0 (and x2 = 1, pure component 2) or x1 = 1 (and x2 = 0, pure component 1).
9
The temperatures at those two points correspond to the boiling points of each of the two pure
components.
For certain pairs of substances, the two curves also coincide at some point strictly between x1 =
0 and x1 = 1. When they meet, they meet tangently; the dew-point temperature always lies above
the boiling-point temperature for a given composition when they are not equal. The meeting
point is called an azeotrope for that particular pair of substances. It is characterized by an
azeotrope temperature and an azeotropic composition, often expressed as a mole fraction. There
can be maximum-boiling azeotropes, where the azeotrope temperature is at a maximum in the
boiling curves, or minimum-boiling azeotropes, where the azeotrope temperature is at a
minimum in the boiling curves.
If one wants to represent a VLE data for a three-component mixture as a boiling point "diagram",
a three-dimensional graph can be used. Two of the dimensions would be used to represent the
composition mole fractions, and the third dimension would be the temperature. Using two
dimensions, the composition can be represented as an equilateral triangle in which each corner
represents one of the pure components. The edges of the triangle represent a mixture of the two
components at each end of the edge. Any point inside the triangle represents the composition of a
mixture of all three components. The mole fraction of each component would correspond to
where a point lies along a line starting at that component's corner and perpendicular to the
opposite edge. The bubble point and dew point data would become curved surfaces inside a
triangular prism, which connect the three boiling points on the vertical temperature "axes". Each
face of this triangular prism would represent a two-dimensional boiling-point diagram for the
corresponding binary mixture. Due to their three-dimensional complexity, such boiling-point
diagrams are rarely seen. Alternatively, the three-dimensional curved surfaces can be represented
on a two-dimensional graph by the use of curved isotherm lines at graduated intervals, similar to
iso-altitude lines on a map. Two sets of such isotherm lines are needed on such a two-
dimensional graph: one set for the bubble point surface and another set for the dew point surface.
The tendency of a given chemical species to partition itself preferentially between liquid and
vapor phases is the Henry's law constant. There can be VLE data for mixtures of four or more
components, but such a boiling-point diagram is hard to show in either tabular or graphical form.
10
For such multi-component mixtures, as well as binary mixtures, the vapor–liquid equilibrium
data are represented in terms of K values (vapor–liquid distribution ratios) defined by
where yi and xi are the mole fractions of component i in the phases y and x respectively.
where is the activity coefficient, Pi is the partial pressure and P is the pressure The values of
the ratio Ki are correlated empirically or theoretically in terms of temperature, pressure and
phase compositions in the form of equations, tables or graph such as the De-Priester charts
(Shown on the right).
11
K-Values for systems of light hydrocarbons Low Temperature Range
12
K-Values for systems of light hydrocarbons High Temperature Range
13
For binary mixtures, the ratio of the K values for the two components is called the relative
volatility denoted by α
which is a measure of the relative ease or difficulty of separating the two components. Large-
scale industrial distillation is rarely undertaken if the relative volatility is less than 1.05 with the
volatile component being i and the less volatile component . K values are widely used in the
design calculations of continuous distillation columns for distilling multi-component mixtures.
For each component in a binary mixture, one could make a vapor–liquid equilibrium diagram.
Such a diagram would graph liquid mole fraction on a horizontal axis and vapor mole fraction on
a vertical axis. In such VLE diagrams, liquid mole fractions for components 1 and 2 can be
represented as x1 and x2 respectively, and vapor mole fractions of the corresponding components
are commonly represented asy1 and y2.[3] Similarly for binary mixtures in these VLE diagrams:
x1 + x2 = 1 and y1 + y2 = 1
Such VLE diagrams are square with a diagonal line running from the (x1 = 0, y1 = 0) corner to
the (x1 = 1, y1 = 1) corner for reference.
These types of VLE diagrams are used in the McCabe–Thiele method to determine the number
of equilibrium stages (or theoretical plates) needed to distill a given composition binary feed
mixture into one distillate fraction and one bottoms fraction. Corrections can also be made to
take into account the incomplete efficiency of each tray in a distillation column when compared
to a theoretical plate.
yA
KA =
xA
where yA = mole fraction of A in the vapor phase and xA = mole fraction of A in the liquid
phase
14
For light hydrocarbon systems (methane to decane), the K values have been determined semi-
emperically and can be evaluated from the equations given in Table 3.1. In general, K is a
function of temperature, pressure, and composition.
yi /xi K
i = = i
yC/xC KC
The values of i will be less dependent on temperature than the values of Ki since the Ki
all increase with temperature in a similar manner.
15
- Choose a component C to be the reference (base) component.
Iteration steps:
2. Evaluate Sumy = • yi
3. Let KC = KC/Sumy
Iteration steps:
2. Evaluate Sumx = • xi
16
3. Let KC = KC Sumx
Solved problems
Let isopentane be the reference compound and T = 582.74oR be a guessed value for the bubble
point calculation (for the above mixture), the next calculated temperature Tcal can be
determined from
Let isopentane be the reference compound and T = 582.74oR be a guessed value for the dew
point calculation (for the above mixture), the next calculated temperature Tcal can be
determined from
17
18
REFERENCES
1. Smith J.M. and Van Ness H.C., Introduction to Chemical Engineering Thermodynamics,
7th Edition, McGraw Hill, 2005.
2. Narayanan K.V., A Text Book of Chemical Engineering Thermodynamics, 3 rd Edition,
Prentice Hall of India Pvt. Ltd.,
2013.
3. Gopinath Halder, Introduction to Chemical Engineering Thermodynamics, 2 nd Edition,
PHI Learning Pvt. Ltd., 2009.
19
SCHOOL OF BIO AND CHEMICAL
DEPARTMENT OF CHEMICAL ENGINEERING
2
3
4.2 Thermodynamic consistency of phase equilibria
➢ The Gibbs/Duhem Equation imposes a constraint on activity coefficient that may not
be satisfied by a set of experimental values derived from P-xy data.
4
•
5
•
6
4.2.2 Using mid point data
7
at
8
4.2.3Redlich/ Kister method
• Diff … w.r.t x1
➢Now ,
9
➢ at
➢ at
➢Plot Vs
10
at const T. by Antion eqn.
11
4.3 Flash distillation of multicomponent mixture using K values
V, yi
F, xiF
Q
L, xi
Defining f = V/F,
12
xiF = fyi + (1 - f)xi
yi = Ki xi = f - 1 xi + xiF
f f
or for xi,
xiF
xi =
f Ki - 1 + 1
The feed composition xiFand the fraction f of the feed vaporized are given at a
specified separator pressure P, the temperature T and compositions xi and yi can be
calculated with the following procedure:
Iteration steps:
2. Evaluate Sumx = • xi at T, P
EndIf
xiF
xi =
f Ki - 1 + 1 and yi = Kixi
13
If the feed composition xiF, temperature T and pressure P of separator are given, then
the fraction of the feed vaporized V/F and compositions xi and yi can be calculated
• yi - • xi = 0
• f KKi -ix1iF+ 1 - • f Ki x- iF1 + 1 = 0
Ki - 1 xiF
•f Ki - 1 + 1
=0
F=
which is known as the Rachford-Rice equation, has excellent convergent properties and can
be solved by Newton’s method. Take the derivative of the function F with respect to V/F (or
f),
dF = - Ki - 1 2xiF
df
• 2
f Ki - 1 + 1
- f = (T - Tb)/(Td - Tb)
Iteration steps:
Ki - 1 xiF
•f Ki - 1 + 1
1. Evaluate F =
dF = - Ki - 1 2xiF
•
2. Evaluate df f Ki - 1 + 1 2
dF
3. Let ER = F/ d f . f = f - ER
xiF
xi =
f Ki - 1 + 1 and yi = Kixi
Solved problems
14
4.1
REFERENCES
15
SCHOOL OF BIO AND CHEMICAL
DEPARTMENT OF CHEMICAL ENGINEERING
Standard free energy changes of formation, Gof, of each species in a change we can
determine the standard state free energy change, Go, for the change using the following
equation:
The equilibrium constant of a chemical reaction is the value of the reaction quotient when the
reaction has reached equilibrium. An equilibrium constant value is independent of the analytical
concentrations of the reactant and product species in a mixture, but depends on temperature and
on ionic strength. Known equilibrium constant values can be used to determine the composition
of a system at equilibrium.
2
Kc is defined as equal to the thermodynamic equilibrium constant but with concentrations of
reactants and products instead of activities. (Kc appears here to have units of concentration raised
to some power while K is dimensionless; however as discussed below under Definitions, the
concentration factors in Kc are properly divided by a standard concentration so that Kc is
dimensionless also.)
Again assuming ideal behavior, the activity of a solvent may be replaced by its mole fraction, or
approximately by 1 in dilute solution. The activity of a pure liquid or solid phase is exactly 1.
The activity of a species in an ideal gas phase may be replaced by its partial pressure.
Stability constants, formation constants, binding constants, association constants and dissociation
constants are all types of equilibrium constant. See also Determination of equilibrium
constants for experimental and computational methods.
Reaction chemistry forms the essence of chemical processes. The very distinctiveness of the
chemical industry lies in its quest for transforming less useful substances to those which are
useful to modern life. The perception of old art of ‘alchemy’ bordered on the magical; perhaps in
today’s world its role in the form of modern chemistry is in no sense any less. Almost everything
that is of use to humans is manufactured through the route of chemical synthesis. Such reactive
processes need to be characterized in terms of the maximum possible yield of the desired product
at any given conditions, starting from the raw materials (i.e., reactants). The theory of chemical
reactions indicates that rates of reactions are generally enhanced by increase of temperature.
However, experience shows that the maximum quantum of conversion of reactants to products
does not increase monotonically. Indeed for a vast majority the maximum conversion reaches a
maximum with respect to reaction temperature and subsequently diminishes.
3
Fiq-5.1 Schematic of Equilibrium Reaction vs. Temperature
The reason behind this phenomenon lies in the molecular processes that occur during a reaction.
Consider a typical reaction of the following form occurring in gas phase:
The reaction typically begins with the reactants being brought together in a reactor. In the initial
phases, molecules of A and B collide and form reactive complexes, which are eventually
converted to the products C and D by means of molecular rearrangement. Clearly then the early
phase of the reaction process is dominated by the presence and depletion of A and B.
However, as the process, continues, the fraction of C and D in the reactor increases, which in
turn enhances the likelihood of these molecules colliding with each other and undergoing
transformation into A and B. Thus, while initially the forward reaction dominates, in time the
backward reaction becomes increasingly significant, which eventually results in the two rates
becoming equal. After this point is reached the concentrations of each species in the reactor
becomes fixed and displays no further propensity to change unless propelled by any externally
imposed “disturbance”(say, by provision of heat).
Under such a condition the reaction is said to be in a state of equilibrium. The magnitude of all
measurable macroscopic variables (T, P and composition) characterizing the reaction remains
constant. Clearly under the equilibrium state the percentage conversion of the reactants to
products must be the maximum possible at the given temperature and pressure. Or else the
4
reaction would progress further until the state of equilibrium is achieved. The principles of
chemical reaction thermodynamics are aimed at the prediction of this equilibrium conversion.
The reason why the equilibrium conversion itself changes with variation of temperature may be
appreciated easily. The rates of the forward and backward reactions both depend on temperature;
however, an increase in temperature will, in general, have different impacts on the rates of each.
Hence the extent of conversion at which they become identical will vary with temperature; this
prompts a change in the equilibrium conversion. Reactions for which the conversion is 100% or
nearly so are termed irreversible, while for those which never attains complete conversion are
essentially reversible in nature.
The fact that a maxima may occur in the conversion behavior suggests that for such reactions
while the forward reaction rates dominate at lower temperatures, while at higher temperatures the
backward reaction may be predominant. The choice of the reaction conditions thus depends on
the maximum (or equilibrium) conversion possible. Further, the knowledge of equilibrium
conversions is essential to intensification of a process. Finally, it also sets the limit that can never
be crossed in practice regardless of the process strategies. This forms a primary input to the
determination of the economic viability of a manufacturing process. If reaction equilibria
considerations suggest that the maximum possible conversion over practical ranges of
temperature is lower than that required for commercial feasibility no further effort is useful in its
further development. On the other hand if the absolute maximum conversion is high then the
question of optimizing the process conditions attain significance. Exploration of the best strategy
for conducting the reaction (in terms of temperature, pressure, rate enhancement by use of
catalytic aids, etc) then offers a critical challenge.
This chapter develops the general thermodynamic relations necessary for prediction of the
equilibrium conversion of reactions. As we shall see, as in the case of phase equilibria, the Gibbs
free energy of a reaction constitutes a fundamental property in the estimation of equilibrium
conversion. The next section presents method of depicting the conversion by the means of the
reaction co-ordinate, which is followed by estimation of the heat effects associated with all
reactions. The principles of reaction equilibria are then developed.
5
5.1 Prediction of free energy data
The Gibbs [free] energy (also known as the Gibbs function) is defined as
In a spontaneous change, Gibbs energy always decreases and never increases. This of course
reflects the fact that the entropy of the world behaves in the exact opposite way (owing to the
negative sign in the TΔS term).
H2O(l)→H2O(s)
water below its freezing point undergoes a decrease in its entropy, but the heat released into the
surroundings more than compensates for this, so the entropy of the world increases, the free
energy of the H2O diminishes, and the process proceeds spontaneously.
An important consequence of the one-way downward path of the free energy is that once it
reaches its minimum possible value, all net change comes to a halt. This, of course, represents
the state of chemical equilibrium. These relations are nicely summarized as follows:
• ΔG = 0: the reaction is at equilibrium; the quantities of [A] and [B] will not change
Finding the value of ΔG: do not bother!
This might seem strange, given the key importance ΔG in determining whether or not a reaction
will take place in a given direction. It turns out, however, that it is almost never necessary to
explicitly evaluate ΔG. As we will show in the lesson that follows this one, it is far more
convenient to work with the equilibrium constant of a reaction, within which ΔG is "hidden".
This is just as well, because for most reactions (those that take place in solutions or gas mixtures)
the value of ΔG depends on the proportions of the various reaction components in the mixture; it
is not a simple sum of the "products minus reactants" type, as is the case with ΔH.
6
Recalling the condition for spontaneous change
ΔG=ΔH–TΔS<0
it is apparent that the temperature dependence of ΔG depends almost entirely on the entropy
change associated with the process. (We say "almost" because the values of ΔH and ΔS are
themselves slightly temperature dependent; both gradually increase with temperature). In
particular, notice that in the above equation the sign of the entropy change determines whether
the reaction becomes more or less spontaneous as the temperature is raised. For any given
reaction, the sign of ΔH can also be positive or negative. This means that there are four
possibilities for the influence that temperature can have on the spontaneity of a process:
As already explained, the above equation implies that if a closed system undergoes a process of
change while being under thermal and mechanical equilibrium, for all incremental changes
associated with the compositions of each species, the total Gibbs free energy of the system would
decrease. At complete equilibrium the equality sign holds; or, in other words, the Gibbs free
energy of the system corresponds to the minimum value possible under the constraints of
constant (and uniform) temperature and pressure. Since the criterion makes no assumptions as to
the nature of the system in terms of the number of species or phases, or if reactions take place
between the species, it may also be applied to determine a specific criterion for a reactive system
under equilibrium.
The general criterion for thermodynamic equilibrium implies that if a closed system undergoes a
process of change while being under thermal and mechanical equilibrium, for all incremental
changes associated with the compositions of each species, the total Gibbs free energy of the
system would decrease. At complete equilibrium the equality sign holds; or, in other words, the
Gibbs free energy of the system corresponds to the minimum value possible under the constraints
of constant (and uniform) temperature and pressure. Since the criterion makes no assumptions as
7
to the nature of the system in terms of the number of species or phases, or if reactions take place
between the species, it may also be applied to determine a specific criterion for a reactive system
under equilibrium. As has been explained in the opening a paragraph of this chapter, at the initial
state of a reaction, when the reactants are brought together a state of non-equilibrium ensues as
reactants begin undergoing progressive transformation to products. However, a state of
equilibrium must finally attain when the rates of forward and backward reactions equalize. Under
such a condition, no further change in the composition of the residual reactants or products
formed occurs. However, if we consider this particular state, we may conclude that while in a
macroscopic sense the system is in a state of static equilibrium, in the microscopic sense there is
dynamic equilibrium as reactants convert to products and vice versa. Thus the system is subject
to minute fluctuations of concentrations of each species.
5.3 The Reaction Coordinate
During the progress of the reaction, at each point the extent of depletion of the reactants, and the
enhancement in the amount of product is exactly in proportion to their respective stoichiometric
th
coefficients. Thus for any change dni in the number of moles of the i species for a differential
Since all terms are equal, they can all be set equal to a single quantity, defined to represent the
extent of reaction as follows: dξ
The general relation between a differential change dnis therefore: (i = 1,2, ...N) iin the number of
moles of a reacting species and dξ
This new variable, called the reaction coordinate, describe the extent of conversion of reactants
to products for a reaction. Thus, it follows that the value of is zero at the start of the reaction. On
the other hand when, it follows that the reaction has progressed to an extent at which point each
8
reactant has depleted by an amount equal to its stoichiometric number of moles while each
product has formed also in an amount equal to its stoichiometric number of moles.
5.4 Nature of Chemical Equilibrium
[B] kf
= = a constant
[A] kr
• this ratio independent of whether reaction started with A or B or a mixture of both (as long as
temp constant)
• forward and reverse reactions do not stop, hence referred to as dynamic equilibrium
• eg. carbonate equilibria (both directions):
• Co2+-complexes:
[Co(H2O)6]2+ + 4 Cl- [CoCl4]2- + 6 H2O
• eg.
H2(g) + I2(g) 2 HI(g)
• eg. start with 1:1 H2:I2 ratio, reaction reaches equilibrium with all 3 components in certain
proportions
• start with different ratio of H2:I2, different proportions of 3 components at equ’m
• start with pure HI, same proportions as first case
9
• i.e.- can approach equ’m from either direction
• quantitative: Law of Mass Action
• general reaction, balanced equation:
aA + bB pP + qQ
• Kc is equilibrium constant (sub “c” denotes concentrations in molarity; sub “p” below)
• numerator for “right side” - prod. of [prod]’s
• denominator for “left side” - prod. of [reactant]’s
• eg. the HI reaction above:
[ HI]2
Kc =
[H 2 ] [I 2 ]
• equilibrium constant depends only on stoichiometry from balanced equation, not on the
reaction mechanism
• K is a true constant at a given temp (varies with temp)
• note: equilibrium constants given without units (above Kc, and Kp below, first developed as
empirical constants; later, thermodynamic equilibrium constants, which are dimensionless)
• when reactants and products are gases, can use pressure instead of molarity
• eg.
N 2O4 (g) 2 NO2 (g)
2
[ NO2 ]2 PNO
Kc = ; Kp = 2
[N 2 O 4 ] PN 2 O 4
10
5.6 Thermodynamic Description of the Equilibrium State
q rev V
but S = and q rev = nRT ln 2 , (ch 7.6)
T V1
V
S = nR ln 2 , (ch 8.5)
V1
P P
= nR ln 1
= - nR ln 2
P2
P1
P P
G = nRT ln 2 = nRT ln
P1 Pref
if choose P1 = 1 atm, the reference state, (then ignore P1 = Pref only if in units of atm)
• Gas-phase reactions
Eg. 3 NO(g) N2O(g) + NO2(g)
= Go + RT ln {(PN2O)(PNO2)/(PNO)3}
In general:
For aA + bB cC + dD
(PC )c (PD )d
Kp =
(PA )a (PB )b
11
Reactions in Ideal Solution
Reactions Involving Pure Solids and Liquids and Multiple Phases; Activity
• if stoichiometric coefficients multiplied, new equ’m constant raised to the power of mult’n
factor:
N2(g) + 3 H2(g) 2 NH3(g)
[ NH3 ]2
K1 = = 3.5 x 108
[N 2 ] [H 2 ]3
12
but for: 1/2 N2(g) + 3/2 H2(g) NH3(g)
[ NH3 ] 1
K2 = 1 3 = K12 = K1 = 1.9 x 104
[N 2 ] [H2 ]
2 2
[N 2 ][H 2 ]3 1
K3 = = K1-1 = = 2.9 x 10-9
[ NH3 ]2 K1
• if two or more equations are added to produce a net equation, the equilibrium constant for the
net equation is the product of the equilibrium constants of the equations added:
• given 3 reactions:
H 2 (g) + Br2 (g) 2 HBr(g); K p = 7.9 x 1011
H 2 (g) 2 H(g); K p = 4.8 x 10-41
Br2 (g) 2 Br(g); K p = 2.2 x 10-15
• using initial concentrations of all components and the equilibrium constant, can also do it
using equilibrium concentrations of some components and the equilibrium constant
13
Magnitude of K and the Direction of Change
• for
1
2 NO2 (g) N 2O4 (g) ; Kc = = 4.72
0.212
[C]c [D]d
Qc =
[A]a [B]b
• add H2 (or N2), system reacts to reduce [H2] towards original value & produce NH3
14
• add NH3, system reacts to reduce [NH3] towards original value & produce H2 + N2
• same conclusion by considering Q:
[ NH 3 ]2
Kc = = Q , at equ' m
[N 2 ] [H 2 ]3
• note: these changes do not change K (const. T), but change Q; re-adjustment until Q = K
15
• when heat added to a system, equilibrium shifts to absorb heat:
• endothermic: increase in T, increased K
• exothermic: increase in T, decreased K
2 NO2(g) N2O4(g); Ho = - 57.2 kJ
brown colorless
• As long as Ho and So not temp-dependent, then T-factor above determines
Extraction and Separation Processes
The phase rule was formulated by J. Willard Gibbs (1874) to determine the variance or
degrees of freedom of a system “f” of a known # of components “c” and phases “p”. The
variance of a system can be expressed as the total variables of this system minus the "fixed
variables". For any system of "c" components and containing "p" phases, the total number of
variables is c.p + 2, where the “2” represents the two variables P and T. The fixed variables
16
are given by the number of equations needed to fully define the composition of the system,
and expressed as: c(p-1) + p. Accordingly, the degrees of freedom will be given by:
f = cp + 2 - [cp - c + p]
f=c-p+2
The phase rule therefore allows us to determine the minimum number of variables that must
be fixed in order to perfectly define a particular condition of the system from a knowledge
of the number of system components and phases. Note that this "number of variables"
cannot be negative (i.e. f 0). The phase rule also allows us to determine the maximum
number of phases that can coexist stably in equilibrium; e.g. if a system has 20 components,
according to the phase rule, the maximum number of phases in equilibrium will be 22 (when
f = 0). Accordingly, if the number of phases present exceeds that calculated by the phase
rule (after the number of components has been correctly identified), then not all these
phases are in equilibrium!
Based on the phase rule, the condition of a system can be described as invariant, univariant,
divariant .....etc., if f = 0, 1, 2,.....etc. respectively, where:
(i) an invariant state is one in which neither P nor T nor X (composition) can be changed
without causing a change in the number of phases present in the system. On the P-T diagram
for the one-component system "SiO2" (which shows the P-T stability fields of the silica
polymorphs; Fig. 1), an invariant state is represented by a point as "C" where three phases
(cristobalite, high quartz and tridymite) coexist.
(ii) a univariant state is one in which either P or T need to be specified in order to fully
define the system. A univariant mineral assemblage can therefore be maintained if a change
in one variable (e.g. P) is accompanied by a dependent change in another (e.g. T), but if one
of these two variables is held constant while the other is changed, the assemblage is no
longer stable, and the system is no longer univariant. An example of this is given by point
"B" of Fig. 1 (or any point lying along any of the curves defining the P-T limits of the
different phases).
17
(iii) a divariant state is one in which two variables have to be specified in order to fully
define or characterize the system. An example of a divariant state is given by point "A" (Fig.
1), or any point lying in the stability fields of one of the polymorphs.
The condensed Phase rule: In cases where either P or T are held constant, one can apply
the "condensed phase rule" given by the formula:
f=c-p+1
This is simply because the total number of variables within the system has now become
pc+1, since only one of the two intensive properties of the system (P and T) is allowed to
vary. The condensed phase rule is quite helpful in understanding isobaric T-X or isothermal
P-X diagrams, and in experimental geochemistry, where either P or T are held constant to
investigate the dependence of the system on the other intensive variable.
In cases where the system consists of dissolved species, defining the number of components
may become challenging. If we decide to consider every dissolved ionic species a
component, then the phase rule is best expressed by the formula:
f = c’ - p –r + 2
where c’ is the number of different chemical species in the system, and r the number of
“auxiliary” restrictions necessary to fully define the system.
To fully appreciate this, we consider the example of a system containing the mineral calcite
at saturation and in equilibrium with gaseous CO2. The system contains 3 phases (water,
calcite, and CO2), and may be considered to conatin 3 components as well (as we attempt to
minimize c in accordance with the definition). Application of the phase rule f = c-p+2 would
indicate that the system is bivariant (3-3+2). However, calcite dissociates in water giving
ionic species, water dissociates to H+ and OH-, whereas CO2 dissolves in water to form
carbonic acid. Possible equilibria in the system therefore include:
18
H2CO3 = H+ + HCO3-
HCO3- = H+ + CO32-
H2O = H+ + OH-
Once we decide to consider the ionic species as system components, c’ = 9 (Ca2+, CO32-,
HCO3-, H+, OH-, H2O, CO2, H2CO3, and CaCO3). “r” then becomes the number of equations
necessary to fully define the system (solve for some concentrations knowing others).
Because c = c’-r, r must be equal to 6. The six equations necessary to fully define the system
become the 5 reactions listed above (one equation for each reaction, each expressing the
equilibrium constant), and an equation that defines the charge balance of the system (all
waters have to be charge – balanced):
19
Solved problems
5.1
20
5.2
21
5.3
5.4
22
23
5.5
REFERENCES
1. Smith J.M. and Van Ness H.C., Introduction to Chemical Engineering Thermodynamics,
7th Edition, McGraw Hill, 2005.
2. Narayanan K.V., A Text Book of Chemical Engineering Thermodynamics, 3 rd Edition,
Prentice Hall of India Pvt. Ltd.,
2013.
3. Gopinath Halder, Introduction to Chemical Engineering Thermodynamics, 2 nd Edition,
PHI Learning Pvt. Ltd., 2009.
24