You are on page 1of 61

The Thermodynamics of

Solutions
Ideal Solutions
• A solution is a homogeneous mixture of two or more components
(substances whose amounts can be independently varied). We
ordinarily apply the name only to solid and liquid mixtures
• Ideal solutions are defined to be solutions in which the chemical
potential of each component is given for all compositions by the
formula

• where μ∗i (T , P) is the chemical potential of the pure component i


when it is at the same temperature, T , and pressure, P, as the
solution and where xi is the mole fraction of component i in the
solution.
Raoult’s Law
• Raoult’s law is an empirical law nearly obeyed by some solutions:

• where Pi is the partial vapor pressure of substance i, defined as the


partial pressure of substance i in the vapor phase that is at
equilibrium with the solution. The equilibrium vapor pressure of the
pure substance i at the temperature and pressure of the solution is
denoted by P∗i , and the mole fraction of substance i in the solution
is denoted by xi.
• From the fundamental fact of phase equilibrium
the chemical potential of component i has the
same value in the solution and in the vapor:

• where we now label the chemical potential of the


gas in its standard state as μ◦(g)i . We will have
additional standard states, so we specify which
one we are using.
• The line segments in the figure correspond to
Raoult’s law, which is very nearly obeyed.
The Thermodynamic Variables of an Ideal
Solution
• The thermodynamic functions of a solution are usually expressed in
terms of the quantities of mixing. These quantities are defined as the
change in the variable for producing the solution from the unmixed
components at the same temperature and pressure.
• The Gibbs energy change of mixing is

• The Gibbs energy change of mixing is

• This equation is identical to that for the Gibbs energy change of


mixing of an ideal gas mixture.
• If an expression for one thermodynamic variable is obtained, the
expressions for other thermodynamic variables can be obtained by
the use of thermodynamic identities. The entropy of a system is given
by

• For the unmixed components, using Euler’s theorem so that

• The enthalpy change of mixing for a solution is given by


Solid Solutions
• In a solid solution the molecules of two or more substances must fit
into a single crystal lattice. Gold and platinum atoms are nearly the
same size and can do this, as can gold and silver atoms.
• These elements are completely miscible and form solid solutions that
are nearly ideal. However, most substances cannot fit into the crystal
lattices of other substances and are nearly insoluble in the solid state,
even if they form nearly ideal liquid solutions.
Phase Diagrams of Two-Component Ideal
Solutions
• The phase diagram for a pure substance requires only two axes
corresponding to T and P. In an equilibrium one-phase system with
two components, Gibbs’ phase rule gives

• so that there are three independent intensive variables, which we


choose to be T and P and one mole fraction. The phase diagram
requires three axes. To make a two dimensional phase diagram, we
must specify a fixed value for one variable.
Pressure–Composition Phase Diagrams
• In this kind of phase diagram the temperature has a
fixed value. The mole fraction of one component is
plotted on the horizontal axis and the pressure is
plotted on the vertical axis.

• where we have used the relationship x2 = 1 − x1. This


equation is represented by a line segment in the
pressure–composition phase diagram.
• A horizontal line segment, or tie line, between the
two curves connects the state points for the two
phases at equilibrium with each other at a given
pressure.
Temperature–Composition Phase
Diagrams
• In this type of phase diagram the pressure is held
fixed. For two components the mole fraction of one
component is plotted on the horizontal axis and the
temperature is plotted on the vertical axis.
• The tie lines drawn between the two curves connect
values of the mole fraction in the two phases at
equilibrium with each other, giving the composition
of one phase as a function of the composition of the
other phase at the temperature given by the height
of the tie line.
• A constant-pressure distillation process can be
described using the temperature– composition phase
diagram
• In order to represent the full equilibrium pressure-temperature-
composition behavior of a two-component system, a three-
dimensional graph is required
Maximum Solubilities in Ideal Solutions
• The mole fraction of naphthalene at which its
chemical potential equals that of the solid
represents the maximum solubility of naphthalene
in a solution with benzene. A solution with this
composition is said to be saturated.
• If a metastable solution occurs with a greater
mole fraction of naphthalene than this, it is called
a supersaturated solution. The maximum solubility
of a solid such as naphthalene in a liquid solvent is
independent of the identity of the liquid solvent if
the two substances form an ideal solution and are
insoluble in the solid phases.
Henry’s Law and Dilute
Nonelectrolyte Solutions
• There is a feature of Figure that is typical of
nonionic substances. For small values of x1 the
curve representing P1 is nearly linear and for
small values of x2 the curve representing P2 is
nearly linear. This behavior corresponds to
Henry’s law, which is written:

• where ki is called the Henry’s law constant for


substance i.
• The Henry’s law constant is not a true constant. It does not depend on
the mole fraction but it depends on temperature and on the identities
of all substances present.
• If one component in a solution is present in a larger amount than the
others, it is called the solvent. We will generally assign the solvent to
be substance number 1. The other substances are called solutes. A
dilute solution is one in which all solutes have small mole fractions.
• In sufficiently dilute solutions nonionic solutes obey Henry’s law and
the solvent obeys Raoult’s law.
The Chemical Potential in a Dilute
Solution
• We write this equation as

• where we let

• The chemical potential μ◦(H)i corresponds to a new standard state for


a solute, which we call the Henry’s law standard state. It is equal to
the chemical potential that the pure liquid would have if it had a
vapor pressure equal to ki instead of P∗i .
Other Measures of Composition
• The molality of component i (a solute) in a solution is defined by

• where ni is the amount of component i in moles and w1 is the mass of the


solvent (component 1) in kilograms. The units of molality are mol kg−1,
sometimes referred to as “molal.”
• For a dilute solution Henry’s law can be expressed in terms of the molality:

• where k(m)i = kiM1 is called the molality Henry’s law constant for
substance i. For a sufficiently dilute solution it is independent of the
molality but depends on the identitiesof all substances present and on the
temperature.
• For a dilute solution, the chemical potential of a solute can be
expressed in terms of the molality

• The quantity μ◦(m)i is the chemical potential of substance i in its


molality standard state. This standard state is component i in a
hypothetical solution with mi equal to m◦ (exactly 1 mol kg−1) and
with Henry’s law valid at this molality.
• The molar concentration of component i is defined by

• If the molar concentration is measured in moles per liter (mol L−1,


sometimes abbreviated as M), it is called the molarity.
Activity and Activity
Coefficients
• We now want to write a single equation that
will apply to all cases:

• where μ◦i is the chemical potential of substance


i in the appropriate standard state and where
this equation defines ai, the activity of
substance i.
• The activity is a dimensionless quantity that is
equal to unity if the substance is in its standard
state.
The Activity of a Pure Solid or Liquid
• The activity of a pure liquid or solid at pressure P is given by

• The exponent in that eq. is generally quite small unless P differs


greatly from P◦.For ordinary pressures we will assume that the activity
of a pure solid or liquid is equal to unity.
The Activity and Activity Coefficient of a
Nonideal Gas
• we define the activity of a nonideal gas to be the ratio of the fugacity
to P◦:

• We define the activity coefficient γi of a nonideal gas as

• The activity coefficent of an ideal gas equals unity. The chemical


potential of a nonideal gas can be written
• The activity coefficient of a gas is also known as the fugacity
coefficient and is sometimes denoted by φi instead of by γi.
• If the value of the activity coefficient of a real gas is greater than
unity, the gas has a greater activity and a greater chemical potential
than if it were ideal at the same temperature and pressure.
• If the value of the activity coefficient is less than unity, the gas has a
lower activity and a lower chemical potential than if it were ideal.
Activities and Activity Coefficients in
Solutions
• The activities are specified in different ways, depending on whether
one of the components is designated as the solvent and depending on
the composition variables used.
• There are two different schemes that use mole fractions as
composition variables, called convention I and convention II.
Convention I
• In this treatment all of the components are treated in the same way,
with no substance designated as the solvent.
• The standard state for each component is the pure substance at
pressure P◦ and at the temperature of the solution, just as with an
ideal solution.
• The chemical potential of substance i in the solution is equal to its
chemical potential in the vapor phase.
• The activity coefficient in convention I is defined as the ratio of the
activity to the mole fraction:
Convention II
• The mole fractions of all components are used as composition
variables in convention II, as in convention I. One substance,
component number 1, is designated as the solvent and is treated just
as in convention I
• The standard state for a solute in convention II is the same as that for
a dilute solute: the hypothetical pure substance that obeys Henry’s
law.
• The chemical potential of the solute in its standard state is equal to
the chemical potential of an ideal gas at pressure equal to ki, which is
the same as the Henry’s law standard state
• The activity is an “effective” mole fraction in expressing the chemical
potential and the partial vapor pressure.

• The activity coefficient in convention II is again defined as the ratio of


the activity to the mole fraction. It is equal to the actual vapor
pressure divided by the vapor pressure predicted by Henry’s law and
describes how the substance deviates from Henry’s law:
The Molality Description
• Activities and activity coefficients of solutes are also defined for the
molality description. The molality of solute number i is given by

• We define the molality activity coefficient

so that
The Concentration Description
• The molar concentration is given by

• We want to write an equation of the form

• so that the activity in the concentration description is

and
The Activities of Nonvolatile
Solutes
The Gibbs–Duhem Integration
• Consider an electrolyte solute represented by the formula Mν+Xν−,
where ν+ and ν− represent the numbers of cations and anions in its
formula. In the case of CaCl2, ν+ 1 and ν− 2. The total number of ions
in the formula is denoted by ν, equal to ν+ + ν−. The substance must
be electrically neutral:

• where z+ is the valence of the cation and z− is the valence of the


anion. The valence is the number of proton charges on an ion. It is
positive for a cation and negative for an anion.
• The (unmeasurable) chemical potential of the cation and that of the
anion are given by

• Since these chemical potentials cannot be measured, we must define


the chemical potential of the neutral electrolyte solute defined by
• In order to write equations that apply to all electrolyte solutes, we
define the mean ionic activity coefficient

• and the mean ionic molality

• For constant pressure and temperature, the Gibbs–Duhem relation for


a two component system is written in the form
• The activity of the solvent can be expressed in terms of the osmotic
coefficient φ, defined by

• where a1 is the activity of the solvent, and M1 is the molar mass of


the solvent. If the solute dissociates completely, νm2 is equal to the
sum of the molalities of the ions.
• It can be shown that φ approaches unity as m2 approaches zero, so
that the integral converges and we can write

• The integral is approximated numerically from the data, giving the


value of γ2 at m2.
The Debye–Huckel Theory
• The principal result of the Debye–Huckel theory is a formula for the
activity coefficient of ions of type i:

• where zi is the valence of the ion (the number of proton charges on the
ion). The quantities α and β are functions of temperature, and of the
properties of the solvent:

• where e is the charge on a proton, NAv is Avogadro’s constant, ρ1 is the


density of the solvent, kB is Boltzmann’s constant (equal to R/NAv), T is
the absolute temperature, and ε is the permittivity of the solvent.
• The quantity I is the ionic strength, defined as a sum over all of the s
different charged species in the solution:

• where mi is the molality of species i and zi is its valence. All of the ions
in the solution must be included in the ionic strength.
• Using the definition of the mean ionic activity coefficient and the
requirement of electrical neutrality,
Thermodynamic Functions of
Nonideal Solutions
• An expression for the partial molar entropy is

• The partial molar enthalpy is given by

• An expression for the partial molar volume is


Thermodynamic Functions of Nonideal
Solutions
• The change in Gibbs energy for producing the solution (the Gibbs
energy change of mixing) is

• This contribution is called the excess Gibbs energy and is denoted by


GE
• The excess entropy can be defined for a nonideal solution:

• The excess enthalpy and the excess volume are equal to the mixing
quantities, since ΔHmix and ΔVmix both vanish for an ideal solution.
• The enthalpy change of mixing is often expressed in terms of the heat
of solution or enthalpy change of solution. For a two-component
solution, the molar integral heat of solution of component 1 in a
solution with component 2 is defined by
• and the molar integral heat of solution of component 2 is defined by

• The differential heat of solution of component 2 is defined by


Phase Diagrams of Nonideal
Mixtures
Liquid–Vapor Phase Diagrams
• This system exhibits positive deviation from Raoult’s law.
• The vapor pressure is larger than it would be if the solution were
ideal, and the solution boils at a lower temperature than if it were an
ideal solution.
• There is such a large positive deviation from
Raoult’s law that there is a maximum in the vapor
pressure curve. the temperature–composition
phase diagram of acetone and chloroform, which
corresponds to a large negative deviation from
Raoult’s law. Either a maximum or minimum
point in a phase diagram is called an azeotrope.
• The two curves representing liquid and vapor
compositions must be tangent at an azeotrope, so
that the two phases have the same composition.
• An azeotropic mixture is sometimes called a
constant-boiling mixture, since it distills without
any change in composition. It is impossible to
distill from one side of an azeotrope to the other.
• Below the temperature labeled Tc there is a
region of tie lines. A tie line in this region
connects points representing the compositions
of the two liquids that can be at equilibrium with
each other. The compositions to the right or to
the left of this region are possible equilibrium
compositions of a single liquid phase.
• The highest point in the tie-line region at
temperature Tc is called an upper critical solution
point or an upper consolute point. It has a
number of properties similar to those of the gas–
liquid critical point
Solid–Liquid Phase Diagrams
• Figure beside shows the solid–liquid temperature–composition phase
diagram of silver and copper at 1.00 atm. There are two one-phase
regions of limited solid solubility, labeled α and β.
• A tie line in the area between the α and β regions represents two
coexisting saturated solid solutions, one that is mostly silver and one
that is mostly copper. The tie line at 779◦C connects the points
representing the two solid phases and one liquid phase that can be at
equilibrium with the two solid phases.
• The point representing this liquid phase is called the eutectic point. If a
liquid that has the same composition as the eutectic is cooled, two
solid phases will freeze out when it reaches the eutectic temperature,
with compositions represented by the ends of the tie line.
Solid–Liquid Phase Diagrams with
Compounds
• Sometimes two substances form solid-state
compounds.
• Figure shows the solid–liquid temperature–
composition phase diagram of aniline (A)
and phenol (P), which exhibit a compound
C6H5NH2 · C6H5OH (abbreviated by AP) in
the solid state
• When it melts an equimolar liquid solution
of aniline and phenol results. This is called
congruent melting.
• Figure 6.22 shows the temperature–composition
phase diagram of copper and lanthanum. There
are four different compounds. There are only two
maxima in the diagram. The two compounds
LaCu and LaCu4 do not melt to form a single
liquid.
• The compound LaCu melts at 551◦C to form two
phases: a liquid solution with lanthanum mole
fraction equal to 0.57 and solid LaCu2. This
phenomenon is called incongruent melting
because the liquid phase does not have the same
composition as the solid phase from which it
formed, and another solid phase is formed.
• The point representing the composition of the
liquid is called a peritectic point. Just as with a
eutectic point, the compositions of the three
phases are not variable at constant pressure.
Three-Component Phase Diagrams
• Figure shows the composition–composition
phase diagram of water, acetone, and ethyl
acetate for a constant temperature of 30◦C and a
constant pressure of 1.00 atm.
• There is a one-phase region of complete
miscibility and a two-phase region containing tie
lines, representing the compositions of two liquid
phases that coexist at equilibrium.
• Since all points in the diagram correspond to the
same temperature and pressure, the tie lines
must remain in the plane of the diagram but are
not required to be parallel to each other.
Colligative Properties
Freezing Point Depression

• where m2 is the molality of the solute. The quantity Kf,1 is called the
freezing point depression constant for the specific solvent.

• The freezing point depression constant Kf,1has a different value for each
solvent, but is independent of the identity of the solute. If there are
several solutes m2 is replaced by the sum of the molalities of all solutes. If
a solute dissociates or ionizes, the total molality of all solute species must
be used, although it might be necessary to account for ion pairing. The
freezing point depression can be used to determine the extent of
ionization of a weak electrolyte.
Boiling Point Elevation

• The boiling point elevation constant for substance 1 is given by

• This quantity has a different value for each solvent, but does not
depend on the identity of the solute. If more than one solute is
present, the molality m2 is replaced by the sum of the molalities of all
solutes.
Vapor Pressure Lowering
• For a nonvolatile solute and a volatile solvent that obeys Raoult’s law,
the total vapor pressure is equal to the vapor pressure of the solvent:

• where P∗1,vap is the vapor pressure of the pure solvent (component


1) and where x1 is the mole fraction of the solvent in the liquid phase.
The lowering of the vapor pressure is given by
Osmotic Pressure
• We denote the pressure on the pure solvent by P, and the pressure on
the solution by P + Π. The difference Π is called the osmotic pressure.
• At equilibrium, the value of the chemical potential of the solvent must
be the same on both sides of the membrane. If the solvent obeys
Raoult’s law,

• where c2 is the molar concentration of the solute. Equation (6.7-23) is


one of two equations known as the van’t Hoff equation. It is
remarkably similar to the ideal gas equation of state.
Exercises

You might also like