Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

CHAPTER 3.

Thermodynamics of Aqueous Phases


Tetsuji Hirato
Department of Energy Science and Technology, Kyoto University, Kyoto, Japan

3.9.1. CHEMICAL POTENTIALS AND


ELECTROCHEMICAL POTENTIALS
The chemical potential of an electrolyte A is given as the partial molar Gibbs
energy, G, as follows


@G
3:9:1
mA
@nA ns ,T ,P
where s refers to the solvent. Electrolytes dissociate into ions (particles with electrical
charges) when they are dissolved in water. For ion species i, the chemical potential
may be defined by
 
@G
3:9:2
mi
@ni nj ,ns ,T ,P
where j refers to all ionic species other than i. However, Equation (3.9.2) does not represent a physically realizable process. It is impossible to add ions of species i only into
solutions, since equal amounts of positive charge on cations and negative charge on
anions are only allowable in water. The energy state of ionic species depends on the
chemical environment. In addition, free energy is required to build up the charge of
the ions. The second energy is clearly proportional to the internal electrical potential,
f, at the phase in question, which can be calculated from electrostatic theory. Although
one cannot experimentally separate these two components for a single species,
Guggenheim [1] separated them conceptually and introduced the electrochemical potential, e
mi , for individual ionic species:
e
mi mi zi Ffi

3:9:3

where zi is the charge number (signed) of the ion and F is the Faraday constant. For reactions in a single conducting phase, fi is constant everywhere in the phase and thus shows
no effect on a chemical equilibrium. The chemical potentials of ions defined by
Treatise on Process Metallurgy, Volume 1
http://dx.doi.org/10.1016/B978-0-08-096986-2.00036-9

2014 Elsevier Ltd.


All rights reserved.

641

642

Tetsuji Hirato

Equation (3.9.2) can be used to calculate chemical equilibriums in aqueous solutions,


although they cannot be measured experimentally. Examples will be shown later in
this chapter.

3.9.2. ACTIVITY AND ACTIVITY COEFFICIENTS


Accepting the use of the chemical potential of a solute in real solutions, the chemical potentials of individual ionic species can be related to their activities, ai:
mi m0i RT ln ai

3:9:4

Since the composition of the solution is described commonly in terms of the molality
scale, usually the hypothetical one-molal solution of the ion is chosen as the standard
state. The standard state is an imaginary solution with molality m01 mol/kg in which
the ions behave ideally. The activity is related to the molality, mi, by
ai gi

mi
m0

3:9:5

where gi is the molal activity coefficient. The activity and activity coefficient have no
dimensions. As the molality approaches zero, obeying Henrys law, gi tends to 1:
gi ! 1 and ai ! mi ! 0

3:9:6

The chemical potential can be written as follows:


RT ln gi
mi m0i RT ln mi RT ln gi mideal
i

3:9:7

mideal
i

is the chemical potential of the ideal-dilute solution of the same molality.


where
Thus, all deviations from ideality are carried in the activity coefficient. The molality scale
can be converted to the mole fraction scale, which is often used in theoretical treatment,
because the mole amount of solvent can be calculated using its formula weight.
For the molarity scale, the hypothetical one-molar solution (c0 1 mol/L) is chosen as
the standard state. The chemical potential can be written using the molarity, ci, and the
molar activity coefficient, yi:
mi m0ic RT ln ci RT ln yi

3:9:8

yi ! 1 and ai ! ci as ci ! 0:

3:9:9

The molarity can be related to the molality by measuring the density of the solution.

3.9.3. MEAN ACTIVITY COEFFICIENTS


The chemical potential of electrolytes as a whole can be measured experimentally.
If the general formula of an electrolyte A is expressed by Mn1 Xn2 , where the cation M and

643

Thermodynamics of Aqueous Phases

the anion X have charges z1e and z2e, respectively, the following equation is obtained
for electrical neutrality:
v1 z1 v2 z2 0

3:9:10

The molar Gibbs energy for a real solution of Mn1 Xn2 is as follows:
mMn1 Xn2 n1 mMz1 n2 mXz2
n1 m0Mz1 n1 RT ln mMz1 n1 RT ln gMz1 n2 m0Xz2
n2 RT ln mXz2 n2 RT ln gXz2

3:9:11

More simply, taking A for the formula of the electrolyte and denoting the cation by 1
and the anion by 2:
mA m0A RT ln mn11 mn22 RT ln gn11 gn22
m0A

n1 m0Mz1

n2 m0Xz2

3:9:12
3:9:13

The mean molarity and mean activity coefficient are defined as follows:
m mn11 mn22 1=n

3:9:14

g gn11 gn22 1=n

3:9:15

and where n n1 n2, Equation (3.9.12) becomes


mA m0A nRT ln m g

3:9:16

Thermodynamic data of electrolytes are often reported and tabulated as the mean
activity coefficient.

3.9.4. THE DEBYEHCKEL LAW [2]


Electrolyte solutions generally consist of solvated ions and solvent molecules. The
departures from ideality in ionic solutions are dominantly caused by the Coulombic
interaction between ions. The contribution of the ionion interaction is estimated on
the basis of the DebyeHuckel theory. Detailed descriptions of the calculation can be
found in specialized books on electrochemistry.
Ions that are oppositely charged attract one another. As a result, anions are more likely
to be found near cations in solution, and vice versa. Near any given ions, there is an excess
of counter ions (ions of opposite charge), even though the overall solution is electrically
neutral. The DebyeHuckel theory relates the activity coefficient of the central ion to its
ionic atmosphere, which is a spherical haze with a net charge equal in magnitude but
opposite in sign to that of the central ion. The electrostatic interaction between the central ion and its ionic atmosphere decreases the energy of the central ion. This energy

644

Tetsuji Hirato

change appears as the difference between the chemical potential and the ideal value of the
solute, and hence can be identified as RT ln g.
In very dilute solutions, the activity coefficient can be calculated from the Debye
Huckel limiting law:
logg jz z jAI 1=2

3:9:17

where A 0.509 for an aqueous solution at 298 K and I is the ionic strength of the
solution:
I

1X 2
mi zi
2 i

3:9:18

The ionic strength is used widely where ionic solutions are discussed.
For more concentrated solutions, where the ionic strength of the solution is too high
(I > 0.2) for the limiting law to be valid, the activity coefficient may be estimated from the
extended DebyeHuckel law:
logg

jz z jAI 1=2


CI
1 BI 1=2

3:9:19

where B and C are other dimensionless constants. Although B can be related to effective
ion radius, it is better considered as an adjustable empirical parameter.

3.9.5. CHEMICAL EQUILIBRIUM AND GIBBS ENERGY


OF FORMATION OF IONS
Consider the following reaction:
aA bB cC dD

3:9:20

The reaction Gibbs energy, DGr, is expressed by using chemical potentials of the
species, A, B, C, and D:
DGr cmC dmD  amA  bmB

3:9:21

By considering Equation (3.9.4), the following formula is obtained:


DGr cm0C dm0D  am0A  bm0B RT ln

acC adD
ac ad
DG0 RT ln Ca D
a b
aA aB
aA abB

3:9:22

where DG0 cm0C dm0D  am0A  bm0B.


Since DGr is zero at equilibrium, one obtains
DG0 RT ln K
where K is the equilibrium constant.

3:9:23

645

Thermodynamics of Aqueous Phases

The value of DG0 can be calculated by using Gibbs energy of formation, DG0f , instead
of chemical potentials, as in the following equation:
DG0 cDGf0 C dDGf0 D  aDGf0 A  bDGf0 B

3:9:24

In aqueous solution, by taking the standard free energy of formation of H(aq) as


zero, the standard free energy of formation of individual ionic species can be obtained.
For example, consider the following reaction:
HClg H aq Cl aq

3:9:25

Here, DGr 0 35:9kJ=mol and DG0f (HCl,g)  95.3 kJ/mol; thus, the standard
free energy of formation of Cl(aq) can be calculated as follows:
DGf0 Cl ,aq 131:2kJ=mol
In the same way, by taking the standard enthalpy of formation and entropy of H(aq)
as zero:
DHf0 H ,aq 0

3:9:26

3:9:27

DS H ,aq 0
0

the standard enthalpy of formation and entropy of individual ionic species can
be obtained.
The data of the standard free energy and enthalpy of formation and entropy of individual ionic species were tabulated by Latimer [3].
The standard free energy of formation at temperature T, DG0T, is given as follows:

DGT0 DH298

T
298

DGT0 DHT0  T DS0T


DCp0 dT  T DS0298  T

T
298

3:9:28
DCp0
T

dT

3:9:29

where DH0, DS0, and DC0p are changes of enthalpy, entropy and specific heat at 1 atm.,
respectively. Knowledge is limited about the specific heat of ionic species in aqueous
solution, so we have to calculate the value of DC0p approximately. First, DC0p is approximated to be zero in the considered temperature range.
0
0
0
0
 TDS298
DG298
 DS298
T  298
DGT0 DH298

3:9:30

Second, if DC0p is approximated to be constant between 298 K and the temperature of


interest, the extrapolation is straightforward.
T
T
1
0
0
0
0
0
dT  T DCp
DGT DH298  TDS298 DCp
dT
3:9:31
298
298 T

646

Tetsuji Hirato

If DH0T is measured at a temperature other than 298 K, the value of DC0p can
be calculated.
As the third approximation, the CrissCobble correspondence principle [4] is commonly applied for simple cations and anions, oxyanions (XOz
n ), and acid oxyanions
z
(HXOn ). The relationship between ionic entropies at 298 K and those at elevated
temperatures can be expressed as follows:
:
ST: aT bT S298

S.T

3:9:32

S.298

where
and
are absolute entropies at the temperature and at 298 K, respectively,
and aT and bT are constants dependent on the temperature and type of ion. The absolute
entropy is related to the conventional entropy (based on S0298 0 for H(aq)) as in the
following equation:

0

S298
i S298
i zS298
H

where z is the ionic charge (signed) and


species, i.

S0298(i)

3:9:33

is the conventional entropy for ion

3.9.6. CHEMICAL EQUILIBRIUM IN AQUEOUS SOLUTIONS


3.9.6.1. AcidBase Reaction
According to the definition by Brnsted and Lowly, acidbase reactions are characterized
as proton transfer reactions. In this system, Brnsted acids donate a proton and Brnsted
bases accept a proton. Here a proton means a hydrogen cation, H:
HAaq H2 Ol A aq H3 O aq

3:9:34

Here, HA(aq) is an acid and H2O(l) is a base, and A(aq) is a conjugate base and
H3O(aq) is a conjugate acid:
Baq H2 O HB aq OH aq

3:9:35

Here, B(aq) is a base and H2O(l) is an acid, HB(aq) is a conjugate acid, and OH(aq)
is a conjugate base.
Although free protons do not exist in aqueous solutions, H(aq) or H is used as a
representation of the state of the proton in this section for simplicity. Accordingly, the
equilibrium constants for reactions (3.9.34) and (3.9.35) are expressed as follows:
aH aA
aHA
aHB aOH
Kb
aB
Ka

where Ka is the acidity constant and Kb is the basicity constant.

3:9:36a
3:9:36b

647

Thermodynamics of Aqueous Phases

The equilibrium constant, KW, for dissociation of water is:


H2 Ol H aq OH aq

3:9:37

3:9:38

KW aH aOH

It is common to use the negative common logarithm of equilibrium constants:


pK logK

3:9:39

Since acidbase reaction is a proton transfer reaction, pH is one of the most important
factors to describe reactions in aqueous solutions, where
pH logaH

3:9:40

In weak acid solutions, pH can be expressed as follows:


pH pKa 

aA
aHA

3:9:41

where the pH value is determined by the ratio of a(A)/a(HA)


pH pKa 

aA
aHA

3:9:42

In aqueous solutions containing metallic ions, hydrolysis reactions must be


considered:
Mn aq nH2 Ol MOHn s nH aq

3:9:43

According to this reaction, the relationship between a(Mn) and pH is determined as


follows:
1
pH fpKMOH  logaMn g
n

3:9:44

where KMOH is the equilibrium constant of the reaction (3.9.43).


In Lewis acidbase reactions, an acid can accept an electron lone pair from another
molecule (Lewis base). Complex formation reaction is one of the Lewis acidbase
reactions:
Mn aq mLaq MLn
m aq

3:9:45

where L represents a ligand. Here the metal ion, Mn(aq), is a Lewis acid and the ligand
is a Lewis base. The equilibrium constants of complex formation reactions have been
tabulated by Martell and Smith [5]. The standard free energy of formation for a complex
species, MLn
m (aq) can be calculated by using its equilibrium constant.

648

Tetsuji Hirato

3.9.6.2. Redox Reactions


Redox reactions (oxidationreduction reactions) are characterized as electron transfer
reactions. Like acidbase reactions, oxidants accept protons and reductants donate electrons. An oxidation reaction does not occur without a reduction reaction. Each reaction
is called a half reaction. In general, a half reaction is presented as follows:
Ox ze Red

3:9:46

where Ox and Red represent an oxidant and a reductant, respectively.


Although the chemical potential of ions cannot be determined as described in
Section 3.9.1, the value of DG0 can be calculated by defining Gibbs energy of formation
of ions. Although the chemical potential of electrons cannot be determined, the equilibrium of reaction (3.9.46) can be considered in the same way, by defining the chemical
potential of electrons as follows:
me FE

3:9:47

where F is the Faraday constant, and E is potential, the potential, E, is the potential based
on the hydrogen standard electrode as described later.
The potential can be obtained on the basis of Equations (3.9.21) and (3.9.22):
E

DG0 RT aRed

ln
zF
zF
aOx

3:9:48

where DG0 DG0f (Red)  DG0f (Ox).


The standard potential, E0, is calculated as follows:
E0 

DG0
zF

3:9:49

By using this Equation (3.9.48) can be expressed as follows:


E E0 

RT aRed
ln
zF
aOx

3:9:50

This equation is called the Nernst equation for half-reactions.


Consider the reduction of hydrogen ions:
2H aq 2e H2 g

3:9:51

where
E0 

DGf0 H2  2DGf0 H
0
2F

3:9:52

649

Thermodynamics of Aqueous Phases

Thus, it is understood that the potential determined by Equation (3.9.47) is the


potential based on the hydrogen standard electrode.
For example, in the Daniell cell,
anodeZnjZnSO4 aqjjCuSO4 aqjCucathode
at the anode, zinc is oxidized as the following half-reaction:
Zn2 aq 2e Zns

3:9:53

At the cathode, copper ions are reduced as the following half-reaction:


Cu2 aq 2e Cus

3:9:54

The total reaction is as follows:


Zns Cu2 aq Zn2 aq Cus

3:9:55

One can obtain the anode and cathode potentials:


RT
aZn
ln
2F aZn2
RT
aCu
ECu2 =Cu E 0 Cu2 =Cu 
ln
2F aCu2
EZn2 =Zn E 0 Zn2 =Zn 

3:9:56
3:9:57

where
DGf0 Zn  DGf0 Zn2
0:76V
2F
DGf0 Cu  DGf0 Cu2
0:34V
E0 Cu2 =Cu 
2F
E0 Zn2 =Zn 

3:9:58
3:9:59

and the electromotive force (emf) of the total reaction can be obtained as follows:
emf ECu2 =Cu  EZn2 =Zn 1:10 

RT aZn2
ln
zF aCu2

3:9:60

3.9.7. POTENTIALpH DIAGRAMS (POURBAIX DIAGRAMS) [6]


Many reactions in aqueous solutions involve the transfer of electrons and/or
protons, which means that redox potentials and the pH of solutions are very important
measures of the tendency of reactions. Therefore, potentialpH diagrams, which was
invented by Marcel Pourbaix (19041998) are useful tools for understanding reactions
in electrolyte solutions. The diagrams are also known as Pourbaix diagrams or
EhpH diagrams.

650

Tetsuji Hirato

In general, reactions in aqueous solutions are represented by the following equation:


mOx nH ze sRed tH2 O

3:9:61

and when z 6 0, one can obtain the potential as a function of pH and activities of species
in aqueous solutions, on the basis of the above discussion.


DG0 2:303RT
aRs

npH
3:9:62

E
log
zF
aOm
zF
When n 0, the reaction is a simple redox reaction.
At 298 K,


DG0 0:0591
aRs
 npH

E
log
zF
aOm
z

3:9:63

When z 0, the reaction is an acidbase reaction and we can obtain the following
relationship:


aRs
0
 npH
3:9:64
DG 2:303RT log
aOm
As DG0  2.303RT log K, the following equation is obtained:
1
aRs 1
logK
pH log
aOm n
n

3:9:65

The stability of water is important in aqueous solutions. Water is decomposed to


H2(g) and O2(g).
2H aq 2e H2 g
E 0:059pH  0:03 logpH2

3:9:66
3:9:67

O2 g 4H aq 4e 2H2 O
E 1:23  0:059pH  0:015logpO2

3:9:68
3:9:69

To obtain Equation (3.9.69), the standard free energy changes of formation of H2O
shown in Table 3.9.1 are used. At potentials higher than Equation (3.9.67) and lower
than Equation (3.9.69), water is stable.
The potentialpH diagram for the ZnH2O system is constructed by using the standard free energy changes of formation shown in Table 3.9.1, as shown in Figure 3.9.1.
The numbered lines in Figure 3.9.1 represent the following reactions:
1,2: Zn2 2e Zn E 0.761 0.0295 log a(Zn2)
3,4: Zn2 2H2O Zn(OH)2 2H, pH 5.86  0.5 log a(Zn2)
5: ZnOH2 HZnO2  H , pH 15:7logaHZnO2 

651

Thermodynamics of Aqueous Phases

Table 3.9.1 Standard Free Energy Change of Formation


Formula
State

DG0f (kJ/mol)

Zn

Zn

aq

147.03

Zn(OH)2

553.58

HZnO
2

aq

464.0

ZnO2
2

aq

389.2

aq

237.2

H2 O

Figure 3.9.1 PotentialpH diagram for the ZnH2O system at 298 K.

6: HZnO2  ZnO2 2 H , pH 13:1 logaZnO =aHZnO


7: ZnOH2 ZnO2 2 2H , pH 14:6 0:5logaZnO
8: Zn(OH)2 2H 2e Zn 2H2O, E 0.416  0.0591pH
9: HZnO2  3H 2e Zn 2H2 O, E 0:05  0:089,pH  0:295logaHZnO
10: ZnO2 2 4H 2e Zn 2H2 O, E 0:44  0:118pH 0:295 logaZnO2
Some may think it strange that reactions concerning Zn2(aq) are considered at
potentials lower than line a, where water is not stable. Strictly speaking, potentialpH
diagrams are not equilibrium phase diagrams. It is, however, useful to consider these reactions in aqueous solutions, since Zn deposition from an aqueous solution actually happens
due to kinetics.

652

Tetsuji Hirato

REFERENCES
[1] E.A. Guggenheim, J. Phys. Chem. 33 (1929) 842, 34, 1540(1930).
[2] P. Atkins, J. de Paula, Atkins Physical Chemistry, eighth ed., Oxford University Press, Oxford, 2006,
pp. 163165.
[3] W.M. Latimer, The Oxidation States of the Elements and Their Potentials in Aqueous Solutions, second
ed., Prentice-Hall, Inc., Englewood Cliffs, N.J., 1952.
[4] J.W. Cobble, et al., J. Am. Chem. Soc. 86 (1964) 5385, 5390, 5396.
[5] A.E. Martell, R.M. Smith, Critical Stability Constants, Plenum Press, New York, NY, 1974.
[6] M. Pourbaix, Atlas of Electrochemical Equilibria in Aqueous Solutions, Pergamon Press, Oxford, 1966.

You might also like