You are on page 1of 111

SIMPLE MIXTURES

Chapter 5
Simple Mixtures
◼ Often in chemistry, we encounter mixtures of
substances that can react together.
◼ Chapter 7 deals with reactions, but let’s first deal
with properties of mixtures that don’t react.
◼ We shall mainly consider binary mixtures –
mixtures of two components.

xA + xB =1
Partial Molar Volumes
◼ Imagine a huge volume of pure water at 25 °C.
If we add 1 mol H2O, the volume increases 18
cm3 (or 18 mL).
◼ So, 18 cm3 mol-1 is the molar volume of pure
water.
Partial Molar Volumes
◼ Now imagine a huge volume of pure ethanol
and add 1 mol of pure H2O it. How much does
the total volume increase by?
Partial Molar Volumes
◼ When 1 mol H2O is added to a large volume of
pure ethanol, the total volume only increases by
~ 14 cm3.
◼ The packing of water in pure water ethanol (i.e.
the result of H-bonding interactions), results in
only an increase of 14 cm3.
Partial Molar Volumes
◼ The quantity 14 cm3 mol-1 is the partial molar
volume of water in pure ethanol.
◼ The partial molar volumes of the components of
a mixture varies with composition as the
molecular interactions varies as the composition
changes from pure A to pure B.
 V 
VJ =  
 n J  p,T ,n'
Partial Molar Volumes
◼ When a mixture is changed by dnA of A and dnB
of B, then the total volume changes by:
 V   V 
dV =   dnA +   dnB
 nA  p,T ,n B  nB  p,T ,n A
Partial Molar Volumes
 V   V 
dV =   dn A +   dn B
 n A  p,T ,n B  n B  p,T ,n A
dV = VA dn A + VB dn B

 
nA nB
V= VA dn A + VB dn B
0 0


Partial Molar Volumes
 
nA nB
V= 0
VA dn A + 0
VB dn B

 
nA nB
V = VA 0
dn A + VB 0
dn B
V = VA n A + VB n B


Partial Molar Volumes
V (NaCl, aq) = volume of sodium chloride solution
_
V (NaCl, aq) = partial molar volume of sodium chloride
in water
Partial Molar Volumes
◼ How to measure partial molar volumes?
◼ Measure dependence of the volume on
composition.
◼ Fit a function to data and determine the slope by
differentiation.
Partial Molar Volumes
◼ Ethanol is added to 1.000 kg of water.
◼ The total volume, as measured by experiment,
fits the following equation:
V = 1002.93 + 54.6664 x − 0.36394 x + 0.028256x
2 3

x = nE
Partial Molar Volumes
 V   V 
VE =   = 
 n E  p,T ,n w  x  p,T ,n w
dV
= 54.6664 − (2)0.36394 x + (3)0.028256x 2

dx
Partial Molar Volumes
◼ Molar volumes are always positive, but partial
molar quantities need not be. The limiting partial
molar volume of MgSO4 in water is -1.4
cm3mol-1, which means that the addition of 1
mol of MgSO4 to a large volume of water results
in a decrease in volume of 1.4 cm3.
Partial Molar Gibbs energies
◼ The concept of partial molar quantities can be
extended to any extensive state function.
◼ For a substance in a mixture, the chemical
potential is defined as the partial molar Gibbs
energy.
 G 
J =  
 n J  p,T ,n'
Partial Molar Gibbs energies
◼ For a pure substance:
G = n J GJ ,m

 G   n J GJ ,m 
J =   =  = GJ ,m
 n J  p,T ,n'  n J  p,T ,n'
Partial Molar Gibbs energies
◼ Using the same arguments for the derivation of
partial molar volumes,
G = nA A + nB B

◼ Assumption: Constant pressure and temperature



Partial Molar Gibbs energies
Fundamental equation of chemical thermodynamics
dG = Vdp − SdT + A dn A + B dn B +

dG = A dn A + B dn B + (at constant p and T)


dw add,max = A dn A + B dn B + (at constant p and T)
Chemical Potential
G = H − TS = U + pV − TS
U = − pV + TS + G
dU = − pdV − Vdp + SdT + TdS + dG
dU = − pdV − Vdp + SdT + TdS + (Vdp − SdT + A dn A + B dn B + )
dU = − pdV + TdS + (A dn A + B dn B + )

dU = A dn A + B dn B + (at constant S and V)


 U 
J =  
 n J  S,V ,n'
Chemical Potential
 H   A 
J =   J =  
 n J S, p,n'  n J V ,T ,n'


Gibbs-Duhem equation
G = n A A + n B B
dG = n A dA + n B dB + A dn A + B dn B

dG = A dn A + B dn B (at constant p and T)

n A dA + n B dB = 0
 n d
J J =0
J
Gibbs-Duhem equation
n A dA + n B dB = 0
nA
dB = − dA
nB


Molarity and Molality
◼ Molarity, c, is the amount of solute divided by
the volume of solution. Units of mol dm-3 or
mol L-1.
◼ Molality, b, is the amount of solute divided by
the mass of solvent. Units of mol kg-1.
Using Gibbs-Duhem
◼ The experimental values of partial molar volume
of K2SO4(aq) at 298 K are found to fit the
expression:
 B = 32.280 + 18.216 x1 2
 B = VK SO
2 4

x = molality of K 2SO 4
Using Gibbs-Duhem
n A dVA + n B dVB = 0
nB
d  A = − d B
nA
nB
 *A d A = −  n d B
A

nB
A −A = − 
*
d B
nA
nB
A =  −
*
A  nA
d B
Using Gibbs-Duhem
nB
A =  −
*
A  nA
d B

 B = 32.280 + 18.216x1 2
d B −1 2
= 9.108x
dx
d B = 9.108x −1 2 dx


Using Gibbs-Duhem
nB
A =  −
*
A  nA
d B

d B = 9.108x −1 2 dx
nB n B −1 2
 A =  A −  9.108x dx =  A − 9.108 
* −1 2 *
x dx
nA nA
Using Gibbs-Duhem
nB n B −1 2
A =  −
*
A  nA
9.108x dx =  A − 9.108 
−1 2 *

nA
x dx

nB nB nB M A
= = = xMA
n A (1 kg) M A 1 kg
 A =  *A − 9.108  xMA x −1 2 dx =  *A − 9.108M A  x1 2 dx


bB
 A =  − 9.108M A
*
A 0
x1 2 dx
 A =  *A − (2 3)9.108M A bB 3 2
Using Gibbs-Duhem
 A =  − (2 3)9.108M A bB
*
A
32

 *A = 18.079 cm3 mol −1


M A = 0.01802 kg mol -1
 A = 18.079 − 0.1094bB 32


Thermodynamics of mixing
◼ So we’ve seen how Gibbs energy of a mixture
depends on composition.
◼ We know at constant temperature and pressure
systems tend towards lower Gibbs energy.
◼ When we combine two ideal gases they mix
spontaneously, so it must correspond to a
decrease in G.
Thermodynamics of mixing
Dalton’s Law
◼ The total pressure is the sum of all the partial
pressure.

pj = x j p
pA + pB + = (x A + x B + )p = p
Thermodynamics of mixing
 p
Gm = Gm + RT ln 
p
 p
 =  + RT ln 
p

 =  + RT ln p


Thermodynamics of mixing
 =  + RT ln p
Gi = n A (A + RT ln p)+ n B (B + RT ln p)
G f = n A (A + RT ln pA )+ n B (B + RT ln pB )
 

pA pB
 mixG = n A RT ln + n B RT ln
p p
Thermodynamics of mixing
pA pB
 mixG = n A RT ln + n B RT ln
p p
pA pA
= xA = xB
p p
 mixG = n A RT ln x A + n B RT ln x B
x A n = nA x B n = nB
 mixG = nRT (x A ln x A + x B ln x B )
 mixG  0
Thermodynamics of mixing
Gibbs energy of mixing
◼ A container is divided into
two equal compartments.
One contains 3.0 mol
H2(g) at 25 °C; the other
contains 1.0 mol N2(g) at
25 °C. Calculate the
Gibbs energy of mixing
when the partition is
removed.
Gibbs energy of mixing
Gi = 3.0(H 2 + RT ln 3p)+ 1.0(N 2 + RT ln p)
G f = 3.0(H 2 + RT ln 3 2 p)+ 1.0(N 2 + RT ln1 2 p)
3
p 2 p
1
 mixG = 3.0(RT ln 2
) + 1.0(RT ln )
3p p
 mixG = 3.0(RT ln 12 ) + 1.0(RT ln 12 )
 mixG = −6.9 kJ

◼ Two processes:
1) Mixing
2) Changing pressures of the gases.
Gibbs energy of mixing

 mixG = nRT (x A ln x A + x B ln x B ) p
3 1
 mixG = 3.0(RT ln ) + 1.0(RT ln )
4 4
 mixG = −2.14 kJ − 3.43 kJ
 mixG = −5.6 kJ

p
Other mixing functions
 G 
  = −S
 T  p,n
  mixG 
 mixS = − 
 T 
 
−  nRT (x A ln x A + x B ln x B ) = −nR(x A ln x A + x B ln x B )
 T 
 mixS = −nR(x A ln x A + x B ln x B )
Other mixing functions
 mixG = nRT (x A ln x A + x B ln x B )
 mixS = −nR(x A ln x A + x B ln x B )

G = H − TS
nRT (x A ln x A + x B ln x B ) = H − T(−nR(x A ln x A + x B ln x B )
nRT (x A ln x A + x B ln x B ) = H + nRT (x A ln x A + x B ln x B )
H = 0 (constant p and T)
Ideal Solutions
◼ To discuss the equilibrium
properties of liquid mixtures we need
to know how the Gibbs energy of a
liquid varies with composition.
◼ We use the fact that, at equilibrium,
the chemical potential of a substance
present as a vapor must be equal to its
chemical potential in the liquid.
Ideal Solutions
◼ Chemical potential of vapor equals the
chemical potential of the liquid at equilibrium.

 =  + RT ln p
*
A A
*
A
*
denotes pure substance
◼ If another substance is added to the pure
liquid, the chemical potential of A will change.

A =  + RT ln pA
A


Ideal Solutions

 = A + RT ln p
*
A
*
A

A =  − RT ln p
*
A
*
A


A = A + RT ln pA
A =  − RT ln p + RT ln pA
*
A
*
A

pA
A =  + RT ln *
*
A
pA
Raoult’s Law
pA = x A p
*
A


Ideal Solutions
pA
A =  + RT ln *
*
A
pA
pA = x A p *
A

A =  + RT ln x A
*
A


Non-Ideal Solutions
Ideal-dilute solutions
◼ Even if there are strong
deviations from ideal
behaviour, Raoult’s law is
obeyed increasingly
closely for the
component in excess as it
approaches purity.
Henry’s law
◼ 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.
Henry’s Law
pB = x B KB
◼ Even if there are strong deviations
from ideal behaviour, Raoult’s law is
obeyed increasingly closely for the
component in excess as it approaches
 purity.
Ideal-dilute solutions
◼ Mixtures for which the solute obeys
Henry’s Law and the solvent obeys
Raoult’s Law are called ideal-dilute
solutions.
Properties of Solutions
◼ We’ve looked at the thermodynamics
of mixing ideal gases, and properties
of ideal and ideal-dilute solutions,
now we shall consider mixing ideal
solutions, and more importantly the
deviations from ideal behavior.
Ideal Solutions
Gi = n A  + n B 
*
A
*
B

A =  + RT ln x A
*
A

G f = n A { + RT ln x A } + n B { + RT ln x B }
*
A
*
B

 mixG = n A RT ln x A + n B RT ln x B
 mixG = nRT{x A ln x A + x B ln x B }
Ideal Solutions

 mixG = nRT{x A ln x A + x B ln x B }

 mixS = −nR{x A ln x A + x B ln x B }

 mixH = 0
Ideal Solutions
 mixG = nRT{x A RT ln x A + x B RT ln x B }

 G 
  =V
 p T

  mixG 
  =  mixV = 0
 p T
Real Solutions
◼ Real solutions are composed of particles
for which A-A, A-B and B-B interactions
are all different.
◼ There may be enthalpy and volume
changes when liquids mix.
◼ G=H-TS
◼ So if H is large and positive or S is
negative, then G may be positive and the
liquids may be immiscible.
Excess Functions
◼ Thermodynamic properties of real
solutions are expressed in terms of excess
functions, XE.
◼ An excess function is the difference
between the observed thermodynamic
function of mixing and the function for an
ideal solution.
Real Solutions

S =  mixS −  mix S
E ideal

G =  mixG −  mixG
E ideal

H =  mixH −  mix H
E ideal
=  mix H
V E =  mixV −  mixV ideal =  mixV


Real Solutions

Benzene/cyclohexane Tetrachloroethene/cyclo
pentane
Colligative Properties
◼ A colligative property is a property that
depends only on the number of solute
particles present, not their identity.
◼ The properties we will look at are: lowering
of vapor pressure; the elevation of boiling
point, the depression of freezing point, and
the osmotic pressure arising from the
presence of a solute.
◼ Only applicable to dilute solutions.
Colligative Properties
◼ All the colligative properties stem from the
reduction of the chemical potential of the
liquid solvent as a result of the presence of
solute.

A =  + RT ln x A
*
A


Thermodynamics of mixing
Boiling Point Elevation
◼ How do we figure out where the new
boiling point is when a solute is present?
◼ Look for the temperature at which at 1
atm, the vapor of pure solvent vapor has
the same chemical potential as the solvent
in the solution.
Boiling Point Elevation
◼ Let’s denote solvent A and solute B.
◼ Equilibrium is established when:

A (g) =  (g) + RT ln x A =  (l) + RT ln x A


*
A
*
A

◼ See Justification 5.1 (Atkins)


*2
RT
T = KxB K=
 vap H
Boiling Point Elevation
*2
RT
T = KxB K=
 vap H

◼ T makes no reference to the identity of


the solute, only to its mole fraction.
 ◼ So the elevation of boiling point is a
colligative property.
Boiling Point Elevation

T = Kx B

For practical purposes :


T = K b b K b = boiling point constant; b = molality
Freezing Point Depression
◼ Let’s denote solvent A and solute B.
◼ Equilibrium is established when:
A (s) =  (l) + RT ln x A
*
A

◼ Same calculation as before (Justification 5.1)


*2
RT
 T = K x B K =
 fusH
Freezing Point Depression

T = K x B

For practical purposes :


T = K f b K f = freezing point constant; b = molality
Cryoscopy

T = K f b K f = freezing point constant; b = molality


n m M
b= =
1 kg solvent 1 kg solvent
Solubility
◼ Although solubility is not strictly a
colligative property (because solubility
varies with the identity of the solute), it
may be estimated using the same
techniques.
◼ When a solid solute is left in contact with a
solvent, it dissolves until the solution is
saturated with the dissolved solute.
Solubility
B (s) =  (l) + RT ln x B
*
B

◼ See Justification 5.2

  fusH  1 1 
ln x B =  − 
R  Tf T 


Osmosis
◼ Osmosis refers to the spontaneous passage
of a pure solvent into a solution separated
from it by a semi-permeable membrane.
◼ In this case, the membrane is permeable to
the solvent but not to the solute.
Osmosis
◼ The osmotic pressure, P, is the pressure
that must be applied to the solution to stop
the influx of solvent.
◼ Examples of osmosis includes the transport
of fluids across cell membranes and
dialysis.
◼ See Justification 5.3
van’t Hoff equation
nB
P = [B]RT [B] =
V
◼ For molecules of large molar mass, such as
polymers or biological macromolecules, a
viral-like expansion used to correct for
 non-ideality. B is the osmotic viral
coefficient.
P = [B]RT{1+ B[B] + ...}
Solvent Activities
◼ The general form of the chemical potential
of a real or ideal solvent is given by:

A =  + RT ln (pA p
*
A
*
A )
◼ For an ideal solution, when the solvent
obeys Raoult’s law, then:

 A =  + RT ln x A
*
A
Solvent Activities
◼ If a solution does not obey Raoult’s law, we
can still use a form of the chemical
potential equation:

A =  + RT ln aA
*
A

◼ The quantity aA is the activity of A. It can


be considered an “effective” mole fraction.

Solvent Activities
◼ Because this equation is true for real or
ideal solvent, we can easily see that:

pA
aA = *
pA


Activities
Calculating solvent activity
◼ The vapor pressure of 0.500 M KNO3(aq)
at 100 °C is 99.95 kPa so the activity of
water in the solution at this temperature is

pA
aA = *
pA
99.95 kPa
aA = = 0.9864
101.325 kPa
Solvent Activities
◼ Because all solvents obey Raoult’s law
increasingly close as the concentration of
solute approaches zero, the activity of the
solvent approaches the mole fraction as
xA=1
pA
*
= xA
pA
aA → x A as x A = 1
Solvent Activities
◼ A way of expressing this convergence is to
introduce the activity coefficient, 

aA →  A x A  A →1 as x A →1
A =  + RT ln x A + RT ln  A
*
A


Solute Activities
◼ Ideal-dilute solutions obeys Henry’s law has
a vapor pressure given by pB=KBxB, where
KB is Henry’s law constant.
pB KB
B =  + RT ln * = B + RT ln * + RT ln x B
*
B
*

pB pB
 KB
B = B + RT ln *
*

pB

B = B + RT ln x B
Solute Activities
◼ For real solutions we can replace xB with aB

 
B = B + RT ln x B = B + RT ln aB
pB
aB =
KB
aB =  B x B
aB → x B and  B →1 as x B → 0
Solute Activities
◼ The selection of a standard state is entirely
arbitrary, so we are free to choose one that
best suits our purpose and description of
the composition of the system

B = B + RT ln bB
bB
aB =  B  where  B →1 as bB → 0
b
B = B + RT ln aB
Activities of regular solutions
◼ Ignore section 5.8
Ion Activities
◼ If the chemical potential of a univalent M+
is denoted + and that of a univalent anion
X- is denoted -, the total molar Gibbs
energy of the ions in the electrically neutral
solution is the sum of these partial molar
quantities.
G ideal
m =ideal
+ +
ideal


Ion Activities
◼ For a real solution of M+ and X- of the
same molality
Gm =  ideal
+ + ideal
− + RT ln  + + RT ln  −
Gm = G ideal
m + RT ln  + −
Ion Activities
◼ Because experimentally a cation cannot
exist in solution without an anion, it is
impossible to separate the product +- into
contributions from each ion, we introduce
the mean activity coefficient

  = ( + − )
12


Ion Activities
◼ The individual chemical potentials of the
ions are:

+ =  ideal
+ + RT ln  
− = −ideal + RT ln  


Ion Activities
◼ If a compound MpXq that dissolves to give
a solution of p cations and q anions.

Gm = p+ + q−
Gm = G ideal
m + pRT ln  + + qRT ln  −
  = (  )
p q 1s
+ − s= p+q
ui = u
ideal
i + RT ln  
Ion Activities
Gm = p+ + q−
ui = u
ideal
i + RT ln  
Gm = G ideal
m + pRT ln   + qRT ln  


Debye-Hückel theory
◼ The departure from ideal behavior in ionic
solutions can be mainly attributed to the
Coulombic interaction between positively
and negatively charged ions.
◼ Oppositely charged ions attract one
another. As a result anions are more likely
found near cations in solution, and vice
versa.
Debye-Hückel theory
Debye-Hückel theory
◼ Although overall the solution is neutrally
charged, but near any given ion there is an
excess of counter ions.
◼ Averaged over time this causes a spherical
haze around the central ion, in which counter
ions outnumbers ions of the same charge as
the central ion, has a net charge the same but
magnitude but opposite sign of the central ion,
and is called the ionic atmosphere.
Debye-Hückel theory
◼ The chemical potential of any given central ion
is lowered as a result of its electrostatic
interaction with its ionic atmosphere.
◼ This lowering of chemical potential is due to
the activity of the solute and can be identified
with RTln±.
Debye-Hückel theory
◼ At very low concentrations the activity
coefficient can be calculated from the
Debye-Hückel limiting law

log   = − z+ z− AI 12

A = 0.509 for an aqueous solution at 25 o


C
zi = charge number of an ion
I = ionic strength of the solution
Debye-Hückel theory
◼ When the ionic strength is too high for the
limiting law to be valid, the activity
coefficient can be estimated from the
Debye-Hückel extended law

12
A z+ z− I
log   = − 12
+ CI
1+ BI


Debye-Hückel theory

1 2  bi 
I =  zi   
2 i b 


Debye-Hückel theory

You might also like