0% found this document useful (0 votes)
58 views25 pages

Chapter 5 - New

This document discusses the thermodynamics of simple mixtures and solutions. It covers: 1) Partial molar quantities like volume which describe how properties change with addition of substances to mixtures. 2) The Gibbs energy of mixing for ideal gas mixtures which is always negative, indicating spontaneous mixing. 3) Properties like entropy of mixing are also discussed for ideal gases. 4) The concept of chemical potential is introduced and shown to describe how extensive properties depend on composition in mixtures and solutions.

Uploaded by

JW M
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
58 views25 pages

Chapter 5 - New

This document discusses the thermodynamics of simple mixtures and solutions. It covers: 1) Partial molar quantities like volume which describe how properties change with addition of substances to mixtures. 2) The Gibbs energy of mixing for ideal gas mixtures which is always negative, indicating spontaneous mixing. 3) Properties like entropy of mixing are also discussed for ideal gases. 4) The concept of chemical potential is introduced and shown to describe how extensive properties depend on composition in mixtures and solutions.

Uploaded by

JW M
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 25

Chapter 5.

Simple mixtures
The thermodynamic description of mixtures
1. Partial molar quantities
Partial molar volume
• For 1 mol H2O added to pure water, the volume increases by 18 cm3
• For 1 mol H2O added to pure ethanol, the volume increases by only
14 cm3
• The partial molar volume: the change in volume per mole of A
added to a large volume of the mixture.
• The partial molar volume, VJ, of a substance J at some general
composition is
𝜕𝑉
𝑉𝐽 = (Definition of partial molar volume)
𝜕𝑛𝐽 𝑝, 𝑇, 𝑛′

where n’ signifies that the amounts of all other substances are constant.
• The partial molar volume is the slope of the plot of the total volume
as the amount of J is changed with p, T, and n’ being held constant.
• When the composition of the mixture is changed by the addition of
dnA of A and dnB of B,
𝜕𝑉 𝜕𝑉
𝑑𝑉 = 𝑑𝑛𝐴 + 𝑑𝑛𝐵 = 𝑉𝐴 𝑑𝑛𝐴 + 𝑉𝐵 𝑑𝑛𝐵
𝜕𝑛𝐴 𝑝,𝑇,𝑛𝐵
𝜕𝑛𝐵 𝑝,𝑇,𝑛𝐴

• Provided the relative composition is held constant as the amounts of


A and B are increased,
𝑛𝐴 𝑛𝐵 𝑛𝐴 𝑛𝐵
𝑉 = න 𝑉𝐴 𝑑𝑛𝐴 + න 𝑉𝐵 𝑑𝑛𝐵 = 𝑉𝐴 න 𝑑𝑛𝐴 + 𝑉𝐵 න 𝑑𝑛𝐵 = 𝑉𝐴 𝑛𝐴 + 𝑉𝐵 𝑛𝐵
0 0 0 0
Partial molar Gibbs energy
• For a substance in a mixture, the chemical potential is defined as
the partial molar Gibbs energy:
𝜕𝐺
𝜇𝐽 = (Definition of chemical potential)
𝜕𝑛𝐽 𝑝, 𝑇, 𝑛′

• For pure substance, G = nJGJ,m and therefore μJ = GJ,m.


• For binary mixture, G = nA μA+ nB μB : The chemical potential of a
substance in a mixture is the contribution of that substance to the
total Gibbs energy of the mixture.
• Because the chemical potentials depend on composition (and the
pressure and temperature), the Gibbs energy of a mixture may
change when these variables change,
(Fundamental equation of
𝑑𝐺 = 𝑉𝑑𝑝 − 𝑆𝑑𝑇 + 𝜇𝐴 𝑑𝑛𝐴 + 𝜇𝐵 𝑑𝑛𝐵 + ⋯
chemical thermodynamics)
• At constant pressure and temperature,
𝑑𝐺 = 𝜇𝐴 𝑑𝑛𝐴 + 𝜇𝐵 𝑑𝑛𝐵 + ⋯

• Under the same condition, dG = dwadd,max:


𝑑𝑤𝑎𝑑𝑑,max = 𝑑𝐺 = 𝜇𝐴 𝑑𝑛𝐴 + 𝜇𝐵 𝑑𝑛𝐵 + ⋯

• That is, additional (non-expansion) work can arise from the changing composition of a
system. For instance, in an electrochemical cell, the chemical reaction is arranged to take
place in two distinct sites (at the two electrodes). The electrical work the cell performs can
be traced to its changing composition as products are formed from reactants.
The wider significance of the chemical potential
𝐺 = 𝐻 − 𝑇𝑆 = 𝑈 + 𝑝𝑉 − 𝑇𝑆 ∴ 𝑈 = −𝑝𝑉 + 𝑇𝑆 + 𝐺
𝑑𝑈 = −𝑝𝑑𝑉 − 𝑉𝑑𝑝 + 𝑆𝑑𝑇 + 𝑇𝑑𝑆 + 𝑑𝐺
= −𝑝𝑑𝑉 − 𝑉𝑑𝑝 + 𝑆𝑑𝑇 + 𝑇𝑑𝑆 + (𝑉𝑑𝑝 − 𝑆𝑑𝑇 + 𝜇𝐴 𝑑𝑛𝐴 + 𝜇𝐵 𝑑𝑛𝐵 + ⋯ )
= −𝑝𝑑𝑉 + 𝑇𝑑𝑆 + 𝜇𝐴 𝑑𝑛𝐴 + 𝜇𝐵 𝑑𝑛𝐵 + ⋯ (Fundamental equation)

𝜕𝑈
• At constant volume and entropy, 𝑑𝑈 = 𝜇𝐴 𝑑𝑛𝐴 + 𝜇𝐵 𝑑𝑛𝐵 + ⋯ ∴ 𝜇𝐽 =
𝜕𝑛𝐽 𝑆, 𝑉, 𝑛′
𝜕𝐻 𝜕𝐴
• In the same ways, 𝜇𝐽 = and 𝜇𝐽 = 𝜕𝑛
𝜕𝑛𝐽 𝑆, 𝑝, 𝑛′ 𝐽 𝑉, 𝑇, 𝑛′

• We see that the μJ shows how all the extensive thermodynamic properties U, H, A, and G
depend on the composition. This is why the chemical potential is so central to chemistry.
The Gibbs-Duhem equation
𝐺 = 𝜇𝐴 𝑛𝐴 + 𝜇𝐵 𝑛𝐵 𝑑𝐺 = 𝜇𝐴 𝑑𝑛𝐴 + 𝜇𝐵 𝑑𝑛𝐵 + 𝑛𝐴 𝑑𝜇𝐴 + 𝑛𝐵 𝑑𝜇𝐵
• At constant pressure and temperature,
𝑑𝐺 = 𝜇𝐴 𝑑𝑛𝐴 + 𝜇𝐵 𝑑𝑛𝐵 𝑛𝐴 𝑑𝜇𝐴 + 𝑛𝐵 𝑑𝜇𝐵 = 0 𝑎𝑛𝑑 ෍ 𝑛𝐽 𝑑𝜇𝐽 = 0 (Gibbs-Duhem equation)
𝐽
• The chemical potential of one component of a mixture cannot change independently of that
of other components. For binary mixture, 𝑑𝜇 = − 𝑛𝐴 𝑑𝜇
𝐵 𝐴
𝑛𝐵
• For all partial molar quantities, if one partial molar quantity increases, then the other must
decrease.
2. The thermodynamics of mixing
The Gibbs energy of mixing of perfect gases
• The chemical potential of pure perfect gas (μ = Gm) is
𝑝 𝑝
𝐺𝑚 (𝑝) = Θ
𝐺𝑚 + 𝑅𝑇 ln Θ ∴ 𝜇 = 𝜇Θ + 𝑅𝑇 ln
𝑝 𝑝Θ
where μo is the standard chemical potential (at 1 bar).
• Let p denote the pressure relative to po : p = p/po.
• Before mixing,
𝐺𝑖 = 𝜇𝐴 𝑛𝐴 + 𝜇𝐵 𝑛𝐵 = 𝑛𝐴 (𝜇𝐴Θ + 𝑅𝑇 ln 𝑝) + 𝑛𝐵 (𝜇𝐵Θ + 𝑅𝑇 ln 𝑝)

• After mixing, the partial pressures are pA and pB, with pA + pB = p.


𝐺𝑓 = 𝑛𝐴 (𝜇𝐴Θ + 𝑅𝑇 ln𝑝𝐴 ) + 𝑛𝐵 (𝜇𝐵Θ + 𝑅𝑇 ln𝑝𝐵 )
• The difference Gf – Gi, the𝑤ℎ𝑒𝑟𝑒
Gibbs 𝑝energy
𝐴 + 𝑝𝐵of
= mixing,
𝑝 ∆mixG, is
𝑝𝐴 𝑝𝐵
Δ𝐺𝑚𝑖𝑥 = 𝑛𝐴 𝑅𝑇 ln + 𝑛𝐵 𝑅𝑇 ln
𝑝 𝑝
• If we use 𝑛𝐽 = 𝑛𝑥𝐽 𝑎𝑛𝑑 𝑥𝐽 = 𝑝𝐽 /𝑝
(Gibbs energy of
𝛥𝐺𝑚𝑖𝑥 = 𝑛𝑅𝑇 𝑥𝐴 𝑙𝑛𝑥𝐴 + 𝑥𝐵 𝑙𝑛𝑥𝐵 mixing of perfect gases)

• Because mole fractions are never greater than 1, the logarithms in


this equation are negative, and ΔmixG < 0: perfect gases mix
spontaneously in all proportions.
Other thermodynamic mixing functions

• Entropy of mixing of perfect gases:

𝜕𝐺 𝜕𝛥𝑚𝑖𝑥 𝐺
= −𝑆 𝛥𝑚𝑖𝑥 𝑆 = − = −𝑛𝑅(𝑥𝐴 𝑙𝑛𝑥𝐴 + 𝑥𝐵 𝑙𝑛𝑥𝐵 )
𝜕𝑇 𝑝,𝑛
𝜕𝑇 𝑝,𝑛𝐴 ,𝑛𝐵

• Because ln x < 0, it follows that ΔmixS > 0 for all compositions.


For equal amounts of gas (xA = xB = 0.5), ΔmixS = nR ln2. This
increase in entropy is what we expect when one gas disperses into
the other and the disorder increases.

• The isothermal, isobaric (constant pressure) enthalpy of mixing of


perfect gases:
𝛥𝐺 = 𝛥𝐻 − 𝑇𝛥𝑆 𝛥𝑚𝑖𝑥 𝐻 = 𝛥𝑚𝑖𝑥 𝐺 + 𝑇𝛥𝑚𝑖𝑥 𝑆 = 0
• The enthalpy of mixing is zero, as we should expect for a system in which there are no
interactions between the molecules forming the gaseous mixture. It follows that the whole
of the driving force for mixing comes from the increase in entropy of the system because
the entropy of the surroundings is unchanged.
3. The chemical potential of liquids
Ideal solutions
• At equilibrium, chemical potential of pure A in vapor and that in
liquid are equal: ∗ Θ ∗
𝜇𝐴 = 𝜇𝐴 + 𝑅𝑇 ln𝑝𝐴 where * denote a quantity related to pure substances
• If another substance is also present in liquid, 𝜇𝐴 = 𝜇𝐴Θ + 𝑅𝑇 ln𝑝𝐴
• Combining the equations, 𝑝𝐴
𝜇𝐴 = 𝜇𝐴∗ − 𝑅𝑇 ln𝑝𝐴∗ + 𝑅𝑇 ln𝑝𝐴 = 𝜇𝐴∗ + 𝑅𝑇 ln
𝑝𝐴∗
• Francois Raoult found that the ratio of the partial vapor pressure of
each component to its vapor pressure as a pure liquid is
approximately equal to the mole fraction in the liquid mixture:
𝑝𝐴 = 𝑥𝐴 𝑝𝐴∗ (Raoult’s law)
• Mixtures that obey the law throughout the composition range from
pure A to pure B are called ideal solutions. Two structurally similar
components may behave almost ideally: ex) benzene and toluene.
• For ideal solution, 𝜇 = 𝜇∗ + 𝑅𝑇 ln𝑥 (Chemical potential of
𝐴 𝐴 𝐴
component of an ideal solution)
• This is also used as the definition of ideal solution.
• Molecular origin of Raoult’s law: In the pure solvent, the vapor pressure represents the
tendency to reach a higher entropy. When a solute is present, the solution has a greater
disorder than the pure solvent. Because the entropy of the solution is higher than that of the
pure solvent, the solution has a lower tendency to acquire an even higher entropy by the
solvent vaporizing.
Ideal-dilute solutions
• For real solutions at low concentrations, although the vapor
pressure of the solute is proportional to its mole fraction, the
constant of proportionality is not the vapor pressure of the pure
substance.
𝑝𝐵 = 𝑥𝐵 𝐾𝐵 (Henry’s law)

where KB is empirical constant.

• Ideal-dilute solution: Mixtures for which the solute obeys


Henry’s law and the solvent obeys Raoult’s law.

• Molecular origin of the difference in behavior of the solute and


solvent at low concentrations: In a dilute solution the solvent
molecules are in an environment very much like the one they
have in the pure liquid, while the solute molecules are
surrounded by solvent molecules, which is entirely different from
their environment when pure. Thus, the solvent behaves like a
slightly modified pure liquid, but the solute behaves entirely
differently from its pure state unless the solvent and solute
molecules happen to be very similar. In the latter case, the solute
also obeys Raoult’s law.
The properties of solutions
4. Liquid mixtures
Ideal solutions
• The total Gibbs energy before liquids are mixed is 𝐺𝑖 = 𝑛𝐴 𝜇𝐴∗ + 𝑛𝐵 𝜇𝐵∗
• The total Gibbs energy after liquids are mixed is
𝐺𝑓 = 𝑛𝐴 𝜇𝐴∗ + 𝑅𝑇ln𝑥𝐴 + 𝑛𝐵 𝜇𝐵∗ + 𝑅𝑇ln𝑥𝐵
• The Gibbs energy of mixing, the difference of these two quantities, is
Δ𝑚𝑖𝑥 𝐺 = 𝑛𝑅𝑇 𝑥𝐴 ln𝑥𝐴 + 𝑥𝐵 ln𝑥𝐵 (Gibbs energy of mixing to form an ideal solution)

Δ𝑚𝑖𝑥 𝑆 = −𝑛𝑅 𝑥𝐴 ln𝑥𝐴 + 𝑥𝐵 ln𝑥𝐵 (Entropy of mixing to form an ideal solution)

Δ𝑚𝑖𝑥 𝐻 = Δ𝑚𝑖𝑥 𝐺 + 𝑇Δ𝑚𝑖𝑥 𝑆 = 0 (Enthalpy of mixing to form an ideal solution)

• The ideal volume of mixing, the change in volume on mixing, is 𝜕Δ𝑚𝑖𝑥 𝐺


Δ𝑚𝑖𝑥 𝑉 = =0
𝜕𝑝 𝑇

• The driving force for mixing is the increasing entropy of the system as the molecules
mingle and the enthalpy of mixing is zero.
• In a perfect gas there are no forces acting between molecules. By contrast, in ideal solutions
there are interactions, but the average energy of A–B interactions in the mixture is the same
as the average energy of A–A and B–B interactions in the pure liquids.
• Real solutions are composed of particles for which A–A, A–B, and B–B interactions are all
different. Not only may there be enthalpy and volume changes when liquids mix, but there
may also be an additional contribution to the entropy arising from the way in which the
molecules of one type might cluster together instead of mingling freely with the others.
Excess functions and regular solutions
• The thermodynamic properties of real solutions are expressed in terms of the excess
functions, XE, the difference between the observed thermodynamic function of mixing and
the function for an ideal solution.
• The excess entropy, SE, is defined as 𝑆 𝐸 = Δ𝑚𝑖𝑥 𝑆 − Δ𝑚𝑖𝑥 𝑆 𝑖𝑑𝑒𝑎𝑙
• The excess enthalpy and volume are both equal to the observed enthalpy
and volume of mixing because ideal values are zero.
• The positive values of HE indicate that the A–B interactions in the mixture
are weaker than the A–A and B–B interactions in the pure liquids (which
are benzene and pure cyclohexane). The symmetrical shape of the curve
reflects the similar strengths of the A–A and B–B interactions.
• Deviations of the excess enthalpy from zero indicate the extent to which
the solutions are nonideal.
• The regular solution: a solution for which HE ≠ 0 but SE = 0; the two
kinds of molecules are distributed randomly (as in an ideal solution) but
have different energies of interactions.
• The excess enthalpy depends on composition as 𝐻𝐸 = 𝑛𝜉𝑅𝑇𝑥𝐴 𝑥𝐵
• ξ is a dimensionless parameter that is a measure of the energy of AB
interactions relative to that of the AA and BB interactions.
• If ξ < 0, mixing is exothermic and the solute–solvent interactions are
more favorable. If ξ > 0, then the mixing is endothermic.
• The Gibbs energy of mixing is 𝛥𝑚𝑖𝑥 𝐺 = 𝑛𝑅𝑇 𝑥𝐴 𝑙𝑛𝑥𝐴 + 𝑥𝐵 𝑙𝑛𝑥𝐵 + 𝜉𝑥𝐴 𝑥𝐵
• For ξ > 2 the graph shows two minima separated by a maximum and the
system will separate spontaneously into two phases.
5. Colligative properties
• The lowering of vapor pressure, the elevation of boiling point, the depression of freezing
point, and the osmotic pressure are arising from the presence of a solute. In dilute solutions
these properties depend only on the number of solute particles present, not their identity. For
this reason, they are called colligative properties (denoting ‘depending on the collection’).
Assumptions : (1) the solute is not volatile
(2) the solute does not dissolve in the solid solvent
The common features of colligative properties
• All the colligative properties stem from the reduction of the
chemical potential of the liquid solvent as a result of the present of
solute.
• Pure solvent → ideal dilute solvent:
𝜇𝐴 = 𝜇𝐴∗ 𝜇𝐴 = 𝜇𝐴∗ + 𝑅𝑇 ln𝑥𝐴

• There is no direct influence of the solute on the chemical potential


of the solvent vapor and the solid solvent because the solute
appears in neither the vapor nor the solid.
• The lowering of the chemical potential occurs even in an ideal solution: its origin is not an
enthalpy but an entropy effect. A solute additionally contribute to the entropy.
• Because of higher entropy of the solution, there is a weaker tendency to form the gas; this
appears as a lowered vapor pressure, and hence a higher boiling point.
• Enhanced randomness of solution opposes the tendency to freeze, lowering the freezing point.
The elevation of boiling point

• The heterogeneous equilibrium is between the solvent vapor and


the solvent in solution at 1 atm:
𝜇𝐴 ∗ (𝑔) − 𝜇𝐴 ∗ (𝑙) ∆𝑣𝑎𝑝 𝐺
𝜇𝐴∗ (𝑔) = 𝜇𝐴∗ (𝑙) + 𝑅𝑇 ln𝑥𝐴 𝑙𝑛 𝑥𝐴 = =
𝑅𝑇 𝑅𝑇
𝑑 𝑙𝑛𝑥𝐴 1 𝑑(∆𝑣𝑎𝑝 𝐺/𝑇) ∆𝑣𝑎𝑝 𝐻 𝜕(𝐺/𝑇) 𝐻
= =− ∵ =− 2
𝑑𝑇 𝑅 𝑑𝑇 𝑅𝑇 2 𝜕𝑇 𝑃
𝑇
𝑙𝑛𝑥𝐴 (Gibbs-Helmholtz equation)
1 𝑇 ∆𝑣𝑎𝑝 𝐻
න 𝑑 𝑙𝑛𝑥𝐴 = − න 𝑑𝑇
0 𝑅 𝑇∗ 𝑇 2
(xA = 1–xB and assume ∆vapH is constant)
∆𝑣𝑎𝑝 𝐻 1 1 ∆𝑣𝑎𝑝 𝐻 ∆𝑇 for xB<<1, ln(1 – xB) = – xB and
ln(1 − 𝑥𝐵 ) = − 𝑥𝐵 =
𝑅 𝑇 𝑇∗ 𝑅 𝑇 ∗2 for T* ~ T, 1/T* – 1/T ~ ∆T/T*2
2
𝑅𝑇 ∗
Δ𝑇 = 𝐾𝑥𝐵 𝑤ℎ𝑒𝑟𝑒 𝐾 =
Δ𝑣𝑎𝑝 𝐻
• No reference to the identity of the solute, only to its mole fraction: the elevation of boiling
point is a colligative property.
• For practical applications, we note that the mole fraction of B is proportional to its molality,
b, in the solution:
Δ𝑇 = 𝐾𝑏 𝑏 (boiling point elevation)

where Kb is the empirical boiling-point constant.


The depression of freezing point
• The heterogeneous equilibrium is between the pure solid solvent A
and the solution with solute present at xB:
2
𝑅𝑇 ∗
𝜇𝐴∗ (𝑠) = 𝜇𝐴∗ (𝑙) + 𝑅𝑇 ln𝑥𝐴 Δ𝑇 = 𝐾 ′ 𝑥𝐵 ′
𝑤ℎ𝑒𝑟𝑒 𝐾 =
Δ𝑓𝑢𝑠 𝐻
Δ𝑇 = 𝐾𝑓 𝑏 (freezing point depression)

where Kf is the empirical freezing-point constant.

The solubility
• Although solubility is not a colligative property, it may be
estimated by the same techniques.
• When a solid solute is left in contact with a solvent, it dissolves
until the solution is saturated; saturation is a state of equilibrium,
with the undissolved solute in equilibrium with the dissolved solute:
𝜇𝐵∗ (𝑠) − 𝜇𝐵∗ (𝑙) 𝛥𝑓𝑢𝑠 𝐺
𝜇𝐵∗ (𝑠) = 𝜇𝐵∗ (𝑙) + 𝑅𝑇 ln𝑥𝐵 𝑙𝑛𝑥𝐵 = =−
𝑅𝑇 𝑅𝑇
Δ𝑓𝑢𝑠 𝐻 1 1
ln𝑥𝐵 = −
𝑅 𝑇𝑓 𝑇
• The solubility of B decreases exponentially as the temperature is lowered from its melting
point. Solutes with high melting points and large enthalpies of melting have low solubilities
at normal temperatures.
• However, it is based on highly questionable approximations, such as the ideality of the
solution.
Osmosis
• The phenomenon of osmosis (from the Greek word for ‘push’) is
the spontaneous passage of a pure solvent into a solution
separated from it by a semipermeable membrane, a membrane
permeable to the solvent but not to the solute.
• The osmotic pressure, ∏, is the pressure that must be applied to
the solution to stop the influx of solvent.
• Important examples of osmosis include transport of fluids through cell membranes, dialysis,
and osmometry, the determination of molar mass by the measurement of osmotic pressure.
Osmometry is widely used to determine the molar masses of macromolecules.
• The chemical potential of the solvent must be the same on each side of the membrane:
𝜇𝐴 ∗ 𝑝 = 𝜇𝐴 (𝑥𝐴 , 𝑝 + Π)
• On the solution side, the chemical potential is lowered by the presence of the solute, which
reduces the mole fraction of the solvent from 1 to xA:
𝑝+Π
∗ ∗
𝜇𝐴 𝑥𝐴 , 𝑝 + Π = 𝜇𝐴 𝑝 + Π + 𝑅𝑇 𝑙𝑛𝑥𝐴 = 𝜇𝐴 𝑝 + න 𝑉𝑚 𝑑𝑝 + 𝑅𝑇 𝑙𝑛𝑥𝐴
𝑝
𝑝+Π
−𝑅𝑇 𝑙𝑛𝑥𝐴 = න 𝑉𝑚 𝑑𝑝 𝑅𝑇𝑥𝐵 = Π𝑉𝑚 Π = [𝐵]𝑅𝑇
𝑝
(van’t Hoff equation)
where xB ~ nB/nA, nAVm= V and [B] = nB/V.
• For non-ideal solutions with large solutes such as proteins and polymers,
𝛱 = [𝐽]𝑅𝑇 1 + 𝐵[𝐽] + ⋯
• The additional terms take the nonideality into account; the empirical constant B is called the
osmotic virial coefficient.
Phase diagrams of binary systems
6. Vapor pressure diagrams
• The partial pressures of the components of an ideal solution of
two volatile liquids are related to the composition of the liquid
mixture by Raoult’s law: 𝑝𝐴 = 𝑥𝐴 𝑝𝐴∗
• The total vapor pressure (at some fixed T) changes linearly with
the composition from pB* to pA* as xA changes from 0 to 1:
𝑝 = 𝑝𝐴 + 𝑝𝐵 = 𝑥𝐴 𝑝𝐴∗ + 𝑥𝐵 𝑝𝐵∗ = 𝑝𝐵∗ + (𝑝𝐴∗ − 𝑝𝐵∗ )𝑥𝐴

The composition of the vapor


• The vapor should be richer in the more volatile component than the liquid.
• From Dalton’s law that the mole fractions in the gas, yA and yB are
𝑝𝐴 𝑝𝐵
𝑦𝐴 = 𝑦𝐵 =
𝑝 𝑝
• For the ideal mixture, 𝑝𝐴 𝑥𝐴 𝑝𝐴 ∗
𝑦𝐴 = = ∗
𝑝 𝑝𝐵 + (𝑝𝐴 ∗ − 𝑝𝐵 ∗ )𝑥𝐴
• For pA*/pB* >1, yA > xA; the vapor is richer than
the liquid in the more volatile component.
• Total pressure can be expressed in terms of yA:
𝑝𝐴 ∗ 𝑝𝐵 ∗
𝑝= ∗
𝑝𝐴 + (𝑝𝐵 ∗ − 𝑝𝐴 ∗ )𝑦𝐴
The interpretation of the diagrams
• The vapor and liquid compositions can be combined to one plot.
• The point a : the vapor pressure of a mixture of composition, xA
• The point b : the composition of the vapor that is in equilibrium
with the liquid at that pressure.
• All the points down to the solid diagonal line is under such high
pressure that it contains only liquid.
• All points below the lower curve correspond to a system that is
under such low pressure that it contains only a vapor phase.
• Points that lie between two lines correspond to where two
phases exist.

• Consider the effect of lowering p on a liquid mixture of overall


composition a. The vertical line with the fixed composition is
“isopleth”
• At a1: the liquid can exist in equilibrium with its vapor. The
composition of the vapor phase is given by point a1’.
• At p1: there is virtually no vapor present; however the tiny
amount of vapor that is present has the composition a1’. (Tie
line: the line that join a1 and a1’)
• At p2: the overall composition is a2’’; the composition of liquid
and vapor phases are a2 and a2’, respectively
• At p3: there is virtually no liquid.
• At a4: only vapor is present; composition is the same as the intial.
The lever rule
• To find the relative amounts of two phases α and β that are in
equilibrium, we measure the distance lα and lβ along the
horizontal tie line and use the lever rule:
𝑛𝛼 𝑙𝛼 = 𝑛𝛽 𝑙𝛽 (Lever rule)
where nα is the amount of phase α and nβ the amount of phase β.
Justification of lever rule:
• Overall amount of A is nzA:
𝑛𝑧𝐴 = 𝑛𝛼 𝑥𝐴 + 𝑛𝛽 𝑦𝐴 and 𝑛𝑧𝐴 = 𝑛𝛼 𝑧𝐴 + 𝑛𝛽 𝑧𝐴 (∵ n = nα + nβ)
𝑛𝛼 ( 𝑧𝐴 − 𝑥𝐴 ) = 𝑛𝛽 ( 𝑦𝐴 − 𝑧𝐴 )
7. Temperature-composition diagrams
The distillation of mixture
• Consider what happens when a liquid of composition a1 is
heated. The liquid boils at T2; the liquid has composition a2 and
the vapor has composition a2’.
• Simple distillation: The vapor is withdrawn and condensed.
• Fractional distillation: The boiling and condensation cycle is
repeated successively. The number of theoretical plates is the
number of effective vaporization and condensation steps that are
required to achieve a condensate of given composition from a
given distillate.
Azeotropes
• A maximum in the phase diagram may occur when the
favorable interactions between A and B molecules reduce the
vapor pressure of the mixture. In such cases, the excess Gibbs
energy, GE, is negative (more favorable to mixing than ideal); ex.
trichloromethane/propanone and nitric acid/water mixtures.
• Consider a liquid of composition a on the right of the maximum.
The vapor (at a2’) of the boiling mixture is richer in A. If the
vapor is removed, the remaining liquid will move to a
composition that is richer in B, a3. As evaporation proceeds, the
remaining liquid shifts towards the composition b.
• Azeotrope: the mixture for which the evaporation occurs
without change of composition. The distillation cannot separate
the two liquids.
• Phase diagrams showing a minimum indicate that the mixture is
destabilized relative to the ideal solution, the A–B interactions
then being unfavorable; GE > 0 and less favorable to mixing
than ideal. There may be contributions from both enthalpy and
entropy effects. Ex. dioxane/water and ethanol/water mixtures.
• Consider a liquid of composition a on the right of the minimum.
The mixture boils at a2 to give a vapor composition a2’. This
vapor condenses to a liquid of a3 with the same composition.
That liquid reaches equilibrium with its vapor at a3’. The
fractionation shifts the vapor towards b, but not beyond.
Immiscible liquids
• At equilibrium, there is a tiny amount of A dissolved in B and a
tiny amount of B dissolved in A. Total vapor pressure is
𝑝 = 𝑝𝐴 ∗ + 𝑝𝐵 ∗
• The ‘mixture’ boils at a lower T than either component would
alone because boiling begins when the total vapor pressure
reaches 1atm, not when either vapor pressure reaches 1 atm.
• Steam distillation enables some heat-sensitive, water-insoluble
organic compounds to be distilled at a lower temperature than
their normal boiling point.
8. Liquid-liquid phase diagrams
Phase separation
• Suppose a small amount of liquid B is added to a sample of
another liquid A at T’. Liquid B dissolves completely forming
a single phase.
• As more B is added, a stage comes at which no more dissolves.
The sample consists of two phases in equilibrium; A-rich
phase (a’’) and B-rich phase (a’); Lever rule is still applied
• As more B is added, a stage comes at so much B is present
and it can dissolve all the A forming a single phase.
• Raising the temperature increases their miscibility; the two-
phase region therefore covers a narrower range of composition.
Critical solution temperatures
• Upper critical solution temperature (UCST), Tuc: the highest
temperature at which phase separation occurs.
• This temperature exists because the greater thermal motion
overcomes any potential energy advantage in molecules of one type
being close together: Ex. nitrobenzene/hexane.
• Gibbs energy of mixing: 𝛥𝑚𝑖𝑥 𝐺 = 𝑛𝑅𝑇 𝑥𝐴 𝑙𝑛𝑥𝐴 + 𝑥𝐵 𝑙𝑛𝑥𝐵 + 𝜉𝑥𝐴 𝑥𝐵
• The phase boundary is at the condition of ∂ΔmixG/∂x = 0
𝑥
𝑙𝑛 + 𝜉(1 − 2𝑥) = 0
1−𝑥
• As ξ decreases, which can be interpreted as an increase in temperature provided the
intermolecular forces remain constant, the two minima move together and merge when ξ = 2.
• Lower critical solution temperature (LCST), Tlc, below which they
mix in all proportions and above which they form two phases: Ex.
Water/triethylamine.
• In this case, at low temperatures the two components are more
miscible because they form a weak complex; at higher temperatures
the complexes break up and the two components are less miscible.
• Some systems have both upper and lower critical solution
temperatures. They occur because, after the weak complexes have
been disrupted, leading to partial miscibility, the thermal motion at
higher temperatures homogenizes the mixture again: Ex. nicotine
and water, which are partially miscible between 61°C and 210°C.
Distillation of partially miscible liquids
• Consider a pair of liquids that are partially miscible and form a low
boiling azeotrope. This combination is quite common because both
properties reflect the tendency of the two kinds of molecule to
avoid each other.
• There are two possibilities: (1) The liquids become fully miscible
before they boil and (2) the boiling occurs before mixing is
complete.
• (1) Distillation of a1 leads to a vapor of composition b1. b1
condenses to the completely miscible single-phase solution at b2.
When this distillate is cooled to b3, phase-separation occurs.
• This description applies only to the first drop of distillate. If
distillation continues, the composition of the remaining liquid
changes. In the end, when the whole sample has evaporated and
condensed, the composition is back to a1.

• (2) Distillate obtained from a liquid of composition of a1 has


composition b3. The distillate is a two-phase mixture of b3’ and b3’’.
• A system at e1 forms two phases, which persist up to the boiling
point at e2. The vapor of this mixture has the same composition as
liquids (azeotrope). Condensing a vapor of composition e3 gives a
two-phase liquid of the same overall composition.
9. Liquid-solid phase diagrams
Eutectics
• a1 → a2: Enters the ‘Liquid+B’, where pure solid B begins to
come out and remaining liquid becomes richer in A.
• a2 → a3: More of the solid B forms.
• a3 → a4: Less liquid than a3 and its composition is e2. This liquid
now freeze to give a two-phase system of pure B and pure A.
• The isopleth e2 corresponds to the eutectic composition, the
mixture with the lowest melting point: a liquid with the eutectic
composition freezes at a single temperature, without previously
depositing solid A or B.
• Solder with of 67% tin and 33% lead by mass melts at 183°C.
• The eutectic formed by 23% NaCl and 77% H2O by mass melts
at −21.1°C; ice can be molten with salt.
• Thermal analysis is useful for detecting eutectics: Liquid cools
steadily from a1 to a2, when B begins to be deposited. Cooling is
now slower because the solidification of B is exothermic and
retards the cooling. When the remaining liquid reaches the
eutectic composition, the temperature remains constant until the
whole sample has solidified: this region of constant temperature
is the eutectic halt. If the liquid has the eutectic composition e
initially, the liquid cools steadily down to the freezing
temperature of the eutectic, when there is a long eutectic halt as
the entire sample solidifies (like the freezing of a pure liquid).
Activities
10. The solvent activities
𝑝
• The chemical potential of solvent is 𝜇𝐴 = 𝜇𝐴∗ + 𝑅𝑇 ln 𝐴∗ 𝜇𝐴 = 𝜇𝐴∗ + 𝑅𝑇 ln𝑥𝐴 (by Raoult’s law)
𝑝𝐴
• For solutions which do not obey Raoult’s law,
𝑝𝐴
𝜇𝐴 = 𝜇𝐴∗ + 𝑅𝑇 ln𝑎𝐴 𝑤ℎ𝑒𝑟𝑒 𝑎𝐴 =
𝑝𝐴 ∗
• The quantity aA is the activity of A, a kind of ‘effective’ mole fraction.
• Because all solvents obey Raoult’s law more closely as the concentration of solute
approaches zero, 𝑎 → 𝑥 𝑎𝑠 𝑥 → 1
𝐴 𝐴 𝐴

• A convenient way of expressing this convergence is to introduce the activity coefficient, γ,


by the definition:
𝑎𝐴 = 𝛾𝐴 𝑥𝐴 𝛾𝐴 → 1 𝑎𝑠 𝑥𝐴 → 1 at all temperatures and pressures
• The chemical potential of the solvent is then
𝜇𝐴 = 𝜇𝐴∗ + 𝑅𝑇 ln𝑥𝐴 + 𝑅𝑇 ln𝛾𝐴
11. The solute activities
Ideal-dilute solutions
• A solute B that satisfies Henry’s law has a vapor pressure given by pB = KBxB.
• In this case, the chemical potential of B is
𝑝𝐵 𝐾𝐵
𝜇𝐵 = 𝜇𝐵∗ + 𝑅𝑇 ln = 𝜇𝐵

+ 𝑅𝑇 ln + 𝑅𝑇 ln𝑥𝐵
𝑝𝐵∗ 𝑝𝐵∗
• Both KB and pB* are characteristics of the solute, so the second term may be combined with
the first to give a new standard chemical potential:
𝐾𝐵
𝜇𝐵Θ = 𝜇𝐵∗ + 𝑅𝑇 ln ∴ 𝜇𝐵 = 𝜇𝐵Θ + 𝑅𝑇 ln𝑥𝐵
𝑝𝐵∗
• If the solution is ideal, 𝐾𝐵 = 𝑝𝐵 ∗ 𝑎𝑛𝑑 𝜇𝐵Θ = 𝜇𝐵∗
• Note that as xB → 0, μB → −∞; that is, as the solution becomes diluted, so the solute
becomes increasingly stabilized and it is very difficult to remove the last traces of a solute.
Real solutes
• Permitting deviations from ideal-dilute (Henry’s law) behavior,
𝑝𝐵
𝜇𝐵 = 𝜇𝐵Θ + 𝑅𝑇 ln𝑎𝐵 𝑤ℎ𝑒𝑟𝑒 𝑎𝐵 = (activity of solute)
𝐾𝐵

• Introducing the activity coefficient for the solute, 𝑎𝐵 = 𝛾𝐵 𝑥𝐵


• Because the solute obeys Henry’s law as its concentration goes to zero,
𝑎𝐵 → 𝑥𝐵 𝑎𝑛𝑑 𝛾𝐵 → 1 𝑎𝑠 𝑥𝐵 → 0 at all temperatures and pressures
• Deviations of the solute from ideality disappear as zero concentration is approached.
12. The activities of regular solutions
• Gibbs energy of mixing to form a non-ideal solution is
Δ𝑚𝑖𝑥 𝐺 = 𝑛𝑅𝑇 𝑥𝐴 ln𝑎𝐴 + 𝑥𝐵 ln𝑎𝐵 = 𝑛𝑅𝑇 𝑥𝐴 ln𝑥𝐴 + 𝑥𝐵 ln𝑥𝐵 + 𝑥𝐴 ln𝛾𝐴 + 𝑥𝐵 ln𝛾𝐵

• Assuming that, ln𝛾𝐴 = 𝜉𝑥𝐵2 𝑎𝑛𝑑 ln𝛾𝐵 = 𝜉𝑥𝐴2 (Margules equations)


Δ𝑚𝑖𝑥 𝐺 = 𝑛𝑅𝑇 𝑥𝐴 ln𝑥𝐴 + 𝑥𝐵 ln𝑥𝐵 + 𝜉𝑥𝐴 𝑥𝐵2 + 𝜉𝑥𝐴2 𝑥𝐵
= 𝑛𝑅𝑇 𝑥𝐴 ln𝑥𝐴 + 𝑥𝐵 ln𝑥𝐵 + 𝜉𝑥𝐴 𝑥𝐵 (𝑥𝐴 + 𝑥𝐵 )
= 𝑛𝑅𝑇 𝑥𝐴 ln𝑥𝐴 + 𝑥𝐵 ln𝑥𝐵 + 𝜉𝑥𝐴 𝑥𝐵
• This is Gibbs energy of mixing. In addition, Margules equations show correct behavior for
𝛾𝐴 → 1 𝑎𝑠 𝑥𝐵 → 0 and 𝛾𝐵 → 1 𝑎𝑠 𝑥𝐴 → 0
• The activity of A is
2 1−𝑥𝐴 2 𝑝𝐴
𝑎𝐴 = 𝛾𝐴 𝑥𝐴 = 𝑥𝐴 𝑒 𝜉𝑥𝐵 = 𝑥𝐴 𝑒 𝜉 and 𝑎𝐴 =
𝑝𝐴∗
1−𝑥𝐴 2
∴ 𝑝𝐴 = 𝑥𝐴 𝑒 𝜉 𝑝𝐴∗

• For ξ = 0, ideal solution and pA=xApA* (Raoult’s law).


• For ξ > 0, endothermic mixing (unfavorable solute-solvent
interaction) gives vapor pressure higher than ideal.
• For ξ < 0, exothermic mixing (favorable solute-solvent
interaction) gives lower vapor pressure.
• All the curves approach linearity as xA → 1: Raoult’s law.
• For xA << 1, 𝑝 = 𝑥 𝑒 𝜉 𝑝∗ = Kx (Henry’s law with K = eξ p *)
𝐴 𝐴 𝐴 A A

You might also like