Sie sind auf Seite 1von 46

CHAPTER 22 PRECIPITATION AND CRYSTALLIZATION KINETICS : MODELS

22.1 Introduction1
22.2 Nucleation... 2
22.2.1 Supersaturation. 2
22.2.2 Effect of Particle Size on Solubility..
3
22.2.3 Nucleation Kinetics...6
22.3 Crystal Growth by Monomer Addition. 10
22.3.1 Growth Models... 10
22.3.2 Monuclear Growth.. 12
22.3.3 Polynuclear Growth19
22.3.4 Spiral Growth..19
22.4 Aggregative Growth
25
22.5 Synthesis of Monodispersed Colloidal Particles... 32
___________________________________________________________________________
22.1 Introduction
Crystallization and precipitation are solubility-related processes. That is, a solid crystal
or precipitate forms when a solute exceeds its solubility in the aqueous solution. In this respect
the solubility product (K
so
) represents a useful parameter for ascertaining the thermodynamic
feasibility of solid-formation reactions. Consider the general reaction:
M
a
A
b
(s) = a M
z+
(aq) + bA
x-
(aq) (22.1)
From a thermodynamic standpoint, the transfer of a solute from the aqueous to the solid phase is
feasible provided the condition log Q > log K
so
is satisfied, where
Q = {M
z+
(aq)}
a
{A
x-
(aq)}
b
(22.2)
and Q = K at equilibrium.
Two different approaches may be taken in the attempt to satisfy the condition of log Q >
log K
so
, i.e., (a) for a given log Q, decrease log K
so
, or (b) for a given log K
so
, increase logQ. In
industrial practice, the term crystallization typically refers to the first case, i.e. the situation
where the condition log Q > log K
so
is achieved by decreasing log K
so
, e.g. via temperature
variation (heating or cooling). On the other hand, precipitation refers to the case where the
condition log Q > log K
so
is attained by altering log Q via the addition of a reagent (precipitant)
which changes {M
z+
} or {A
x-
}. This chapter is devoted to a discussion of the kinetics of
precipitation and crystallization processes.
22.2 Nucleation
22.2.1 Supersaturation
A precipitation or crystallization process involves three main steps, i.e., the development of
supersaturation, followed by nucleation, and then growth. Supersaturation may be defined in
several ways, e.g., as the supersaturation ratio (S), the concentration driving force (C), and the
relative supersaturation (). The relevant relations are given by Equations 22.3, 22.4, and 22.5:
S = C/C
eq
(22.3)
C = C - C
eq
(22.4)
= C/C
eq
= S - 1 (22.5)
where C and C
eq
respectively represent the solute concentrations in the supersaturated and
equilibrium saturated solutions.
The precipitation of relatively insoluble compounds exhibits many of the characteristics
associated with the crystallization of relatively soluble salts (i.e., salts with C
eq
> 0.1 to 1 kmol
m
-3
). However, precipitation differs from crystallization in that it involves a much higher
degree of supersaturation (S = (C-C
eq
)/C
eq
~ 1,000 for precipitation compared with ~ 0.001 to
0.20 for crystallization systems). As a consequence of this comparatively high degree of
supersaturation, the tendency in precipitation processes is to form relatively small crystals.
The supersaturation ratio (S) represents the thermodynamic driving force for
crystallization. Consider the precipitation reaction,
M
z+
+ A
z-
= MA(s) (22.6)
The corresponding Gibbs free energy of reaction is given by
G
r
= -RT ln K + RT ln Q (22.7)
where K is the equilibrium constant:
K = 1/{M
z+
}
eq
{A
z-
}
eq
(22.8)
and Q is the reaction quotient:
Q = 1/{M
z+
} {A
z-
} (22.9)
It follows from Equations 22.7-22.9 that
G
r
= - RT ln (C/C
eq
) (22.10)
= - RT lnS (22.11)
The precipitation reaction will be favored, i.e., Equation 22.6 will proceed in the forward
direction if G
r
< 0. Therefore on the basis of Equations 22.10 and 22.11, precipitation will
proceed so long as C > C
eq
or S > 1

EXAMPLE 22.1. Thermodynamic Driving Force for Crystallization

22.2.2 Effect of Particle Size on Solubility
In preparation for our later discussion of nucleation kinetics, it is helpful to consider
particle size effects on solubility. Consider a solid particle bounded by an arbitrary
solid/aqueous interface. The Gibbs free energy is given by:
dG = -SdT + Vdp + dA +
i
i
i
dn
(22.12)
where S = entropy, T = absolute temperature, V = volume, p = pressure, = interfacial energy, A
= interfacial area,
i
= chemical potential of component i, n
i
= moles of component i. The
chemical potential is given by

i
= ( G/ n
i
)
T,P,nj,A
(22.13)
Let us focus on a solid consisting of only one type of compound, i.e., the solid contains
only one component. Then at constant temperature and pressure, Equation 22.13 becomes:
dG = dA + dn (22.14)
where is the bulk chemical potential of the crystal.
Equation 22.13 assumes that n
i
can change at constant A. For very small particles,
however, n cannot be transferred without changing A. If dV corresponds to the volume change
associated with introduction of dn molecules, then it follows that
dV =
m
dn (22.15)
where
m
is the molar volume of the solid particle. But A = 4r
2
and V = (4/3)r
3
and thus
dA = 8rdr = 2dV/r (22.16)
Therefore it follows from Equations 22.15 and 22.16 that
dA = 2
m
dn/r (22.17)
Introducing Equation 22.17 into Equation 22.14,
dG = (2
m
/r + ) dn (22.18)
Thus ' the chemical potential of the microparticle is given by
' = ( G/ n)
T,p,nj
= 2
m
/r + (22.19)
Comparing Equations 22.13 and 22.19 it can be seen that in the case of the microparticle, the
constant A constraint has been removed.
Consider the case where a solute distributes between the aqueous phase and the solid
particle. If ' and are respectively the chemical potentials of the solute for a microparticle and
for macroparticle, then it follows from Equation 22.19 that
' - = 2
m
/r (22.20)
Now the chemical potential of the solute in the aqueous phase in contact with the macroparticle
is given by:
= + RT ln a (22.21)
where a is the aqueous phase activity of the solute. Similarly for the case of the microparticle,
' = + RT ln a' (22.22)
At equilibrium the chemical potentials of the solute in the contacting solid and aqueous
phases are equal. Therefore,

s
= = + RT ln a (22.23)

'
s
= ' = + RT ln a' (22.24)
Therefore combining Equation 22.20, 22.23, and 22.24 gives
ln (a'/a) = 2
m
/rRT (22.25)
Under conditions where activity coefficient effects can be ignored, Equation 22.25 may be
expressed in terms of the solute concentrations as:
ln (C'/C) = 2
m
/rRT (22.26)
= 2M/rRT (22.27)
where M is the molecular weight of the solid compound and is the density of the solid.
______
EXAMPLE 22.2. Effects of particle size on solubility and crystal growth.
(a) Show that the particle size effect on solubility can be incorporated into the expression for the solubility
product constant as:
log K
so
(r) = log K
so
() + (2/3) A/2.3RT (1)
where K
so
(r) represents the solubility product for a small precipitate with a molar surface area
A(meter
2
/mole), K
so
() is the solubility product constant for a large precipitate.
(b) On the basis of Equation 1 and the data provided below, derive expressions for the solubility product
constants of Cu(OH)
2
(s) and CuO(s) which take surface effects into account.
Reaction log K
so
( ) (erg cm
-2
)
CuO(s) + H
2
O = Cu
2+
+ 2OH
-
-20.0 690
Cu(OH)
2
(s) = Cu
2+
+ 2OH
-
-19.0 410

(c) Consider an aqueous solution at pH 7.0. As the Cu
2+
concentration is increased, which of the two solids
Cu(OH)
2
and CuO will nucleate first? Assume that the precipitate has an initial molar surface area of
5x10
4
m
2
/mole.
(d) As crystal growth proceeds, the compound which first precipitates becomes unstable relative to the
second compound. At what value of molar surface area does the first precipitate invert to give the
second solid? What is the equilibrium Cu
2+
concentration at this stage?
____________________________________________________________________________________________
22.2.3 Nucleation Kinetics
The establishment of supersaturation does not lead automatically to solid formation. For
this, it is necessary that crystal nuclei be formed. In general, the ions present in solution are in
constant motion and their mutual interaction results in the formation of temporary clusters. The
process leading to the formation of a permanent or critical cluster, i.e., a crystal nucleus, can be
viewed as a succession of bimolecular reactions:
M + M M
2
M
2
+ M M
3
M
n-1
+ M M
n
(critical cluster) (22.28)
In view of its extremely small size, it is practically impossible to determine the structure of this
critical cluster. However, it is apparent that clusters which contain less molecules than the
critical cluster are unstable and redissolve.
It can be shown from a thermodynamic analysis of the free energy changes associated
with the nucleation process that a cluster with a radius greater than that of the critical cluster (r
c
)
is stable because if it grows, it can reduce its free energy. On the other hand, a molecular
assemblage with a radius less than r
c
can only decrease its free energy by decreasing in size, i.e.,
dissolving.
The free energy change, G
n
, associated with nucleation has two main contributions.
When molecules or ions assemble into a cluster, energy is given off (e.g. latent heat of

condensation when gaseous molecules condense to liquid). The corresponding free energy
change, G
v
, is therefore a negative quantity. On the other hand, creation of the aggregate
results in the formation of a new surface. The corresponding interfacial energy, G
s
, is a
positive quantity. Thus
G
n
= G
v
+ G
s
(22.29)
= (4/3)r
3
G

+ 4r
2
(22.30)
where is the interfacial energy and G

is the free energy of aggregation per unit volume of


aggregate.
G
crit
r
crit
r

G
s
G
n
G
v
G
0
Figure 22.1 The effect of the cluster radius (r) on the free energy of nucleation (G
n
).
The radius of the critical nucleus is obtained when dG
n
/dr = 0, i.e.,
dG
n
/dr = 8r

+ 4r
2
G

= 0 (22.31)

Thus,
r
c
= -2/G

(22.32)
Inserting Equation 22.32 into Equation 22.30 gives:
G
crit
= 16
3
/3(G

)
2
= 4r/3 (22.33)
Figure 22.1 presents a schematic illustration of the dependence of r on G
n
. From the free
energy diagram it follows that a cluster with a radius greater than r
c
is stable because if it grows
it can reduce its free energy (i.e. |G

| > G
s
). On the other hand, a particle with a radius less
than r
c
can only decrease its free energy by decreasing in size, i.e. dissolving.
How is the amount of energy needed to overcome the energy barrier, G
crit
, obtained?
At constant temperature and pressure, the average energy in the system is constant. However the
energy fluctuates around this average value and it is therefore possible for some parts of the fluid
to momentarily acquire energy levels which will permit the corresponding clusters to overcome
G
crit
.
The transition from a critical radius to a stable nucleus can be viewed in terms of the
transformation of a microparticle to a macroparticle, as discussed in Section 22.2.2.
Accordingly, the solute concentration (C') associated with the microparticle represents the solute
concentration in the supersaturated solution that surrounds the nucleus. On the other hand the
solute concentration (C) in the solution adjacent to the macroparticle may be linked to the solute
concentration in the saturated solution obtained at equilibrium around the fully grown particle.
Thus recalling Equations 22.3 and 22.26,
ln (C'/C) = ln S = 2
m
/r
c
RT (22.34)
Therefore
r
c
= 2
m
/RT ln S (22.35)
Combining Equation 22.33 with Equation 22.35 gives
G
crit
= 16
3
/3 (RT ln S)
2
(22.36)

An Arrhenius type of expression can be used to relate the rate of nucleation and G
crit
:
R
n
= k
o
exp(-G
crit
/RT) (22.37)
where k
o
is a constant. It follows from Equations 22.36 and 22.37 that
R
n
= k
o
exp[-16
3
/3R
3
T
3
(ln S)
2
] (22.38)
According to Equation 22.38, three key variables control the rate of nucleation:
temperature (T), interfacial energy (), and the degree of supersaturation (S). In particular, it can
be seen from Equation 22.38 that as the supersaturation increases, the nucleation rate also
increases.
When nucleation proceeds without assistance from foreign substances, including the
container walls, the process is termed homogeneous nucleation. On the other hand when
external agents such as seed crystals or container walls are used to initiate nucleation, the process
is termed heterogeneous nucleation. The ability of solid surfaces to catalyze nucleation reactions
stems from the fact that the activation energy for the formation of a two-dimensional nucleus is
smaller than that required for a three-dimensional nucleus.
___________________________________________________________________________
EXAMPLE 22.3 The number of moles in a critical nucleus
(a) Starting from Equation 22.35, show that n
c
, the number of moles in a critical nucleus is given by
n
c
= 32
3
/3(RT ln S)
3
(1)
(b) Under certain conditions the experimentally determined nucleation rate is found to follow the
following power law:
R = k C
n
(2)
Show that n = n
c
.
Solution
(a) According to Equation 22.35,
r
c
= 2
m
/RT ln S
where
m
is the molar volume of the solute. Let the volume of the critical nucleus be V
c
. Then the
number of moles in the critical nucleus is given by
n
c
= V
c
/
m
= (4r/3)/
m
= 32
3
/3(RT ln S)
3
(3)
(b) It follows from Equation 2 that:
ln R = lnk + n ln C (4)
Thus
n = d ln R/d ln C (5)
But according to Equation 22.38,
R = k
o
exp[-16
3
/3 R
3
T
3
(ln S)
2
] (22.38)
Thus
ln R = ln k
o
- 16
3
/3 R
3
T
3
(ln S)
2
(6)
Therefore
= (7)
Comparison of Equations 1 and 7 indicates that
d ln R/d ln S = n
c
(8)
But by definition, S = C/C
eq
(see Equation 22.3). Therefore
d ln R/d ln S = d ln R/d ln C (9)
It follows from Equations 5, 8, and 9 that n = n
c

22.3 Crystal Growth by Monomer Addition
22.3.1 Growth Models
Once a stable nucleus is formed, it can grow by two main pathways, i.e., monomer
addition, and nuclei aggregation, as illustrated in Figure 22.2.
In the past, the formation of primary particles was viewed mostly in terms of monomer
addition. However, there is increasing evidence that in some cases (e.g. oxide and sulfide
precipitations, formation of colloidal gold) aggregative growth constitutes a significant pathway
in the formation of primary particles.
PRECURSOR
INTERMEDIATE (MONOMER)
NUCLEI (SUBPARTICLES)
PRIMARY PARTICLES PRIMARY PARTICLES
AGGREGATES
+ PRIMARY
PARTICLES
+ PRIMARY
PARTICLES
+ MONOMERS + NUCLEI
AGGREGATIVE
GROWTH
MONOMER
ADDITION
GROWTH
Figure 22.2 Stages in the particle growth process
___________________________________________________________________________
EXAMPLE 22.4 Aggregation of silica nanoparticles
(a) According to the Smoluchowski theory of rapid aggregation (i.e., where there is no energy barrier) the time
(t
1/2
) taken for an initial particle number density of C
o
to decrease by one-half is given by
t
1/2
= 3/4k
B
TC
o
(1)
where = viscosity of the solvent, k
B
= Boltzmann constant, T = absolute temperature. An experiment to
synthesize silica particles (by the ammonia-catalyzed hydrolysis of alkoxides) yields a number density of
10
18
m
-3
. Determine the half-life for aggregation, given that the viscosity of the continuous phase is 10
-3
Pa s.
(b) In the presence of an energy barrier, Equation 1 may be modified as:
t
1/2
= 3W
ii
/4k
B
TC
o
(2)
where W
ii
is termed the "stability ratio for two particles of size i". In the case of equal-sized particles
where the interaction forces are derived solely from van der Waals attraction and electrostatic repulsion,
W
ii
is given approximately by
W
ii
= (1/2 a)exp [4
o

o
2
/k
B
T] (3)
where
= relative permitivity of the liquid

o
= permitivity of free space

o
= surface potential of the particle
a = particle radius
= reciprocal of the double-layer thickness (see Chapter 5).
Assuming a surface potential of 25mV, determine the half-life for particles with a radius of (i) 25nm,
(ii) 5nm.
Solution
(a) It can be shown that the relevant half-life is 0.1s.
(b) It can be shown that for the 25nm particles, t
1/2
> 12 h whereas for the 5 nm particles, t
1/2
< 1s.
____________________________________________________________________________________________
Example 22.4a illustrates the fact that in the absence of an energy barrier, aggregation is
highly favored. Thus the formation of a colloidally stable precipitate implies the presence of a
sufficiently high energy barrier (e.g. electrostatic, solvation, steric). On the other hand Ex. 22.4b
illustrates the fact that even in the presence of electrostatic repulsion, extremely small particles
(approaching the size of a nucleus) are unstable towards aggregation.
22.3.2 Mononuclear Growth
As illustrated in Figure 22.3, the growth of an ionic crystal by monomer addition may be
viewed in simple terms as a two-step process involving (a) the diffusion of ions from the bulk
solution through a stagnant boundary layer to an adsorption layer:
J
d
= (D/)(C - C
i
) (22.39a)
= k
d
(C - C
i
) (22.39b)

SOLID
BULK AQUEOUS
SOLUTION

C
C
i
(D/ ) C
i
(C- ) J
d
=
ADSORPTION
LAYER
AQUEOUS BOUNDARY
LAYER
Figure 22.3 Crystal growth via combined film diffusion and surface reaction.
and (b) interfacial reaction, i.e. the incorporation of the adsorbed ions into the crystal lattice:
M(ads) M(lattice) (22.40)
Assuming a first-order reversible reaction at the surface, the flux (J
r
) associated with the
reaction can be expressed as:
J
r
= (1/A) dn/dt = k
1
C
i
- k
-1
C
L
(22.41a)
= k
1
C
i
- k (22.41b)
where n is the moles of the solid present at time t, A is the solid/solution interfacial area, and C
L
is the concentration of the ion in the crystal lattice (which is a constant). At equilibrium dn/dt = 0
and C
i
= C
eq
and therefore it follows from Equation 22.41 that:
k
1
C
eq
= k (22.42)
and
J
r
= k
1
(C
i
- C
eq
) (22.43)
At steady-state the rate of diffusion must equal the rate of the surface reaction. Thus it
follows from Equations 22.39 and 22.43 that
J = J
d
= J
r
= k(C - C
eq
) (22.44)
where k is the overall mass transfer coefficient for crystal growth given by,
1/k = 1/k
d
+ 1/k
1
(22.45)
It can be seen from Equations 22.44 and 22.45 that when the surface reaction is very fast, k~k
d
and crystal growth is diffusion controlled. On the other hand, when diffusion is relatively fast,
k~k
r
and crystal growth is surface reaction controlled.
It can be seen from Equation 22.44, that for a first-order surface reaction, the crystal
growth rate is proportional to the concentration difference (C-C
eq
). It must be noted that here
C
eq
refers to the solute concentration in equilibrium with the surface of the freshly precipitated
solid. Thus if the surface of the fresh precipitate eventually undergoes rearrangement into a
more stable form, the solute concentration (C
o
eq
) in equilibrium with this more stable form may
differ from C
eq
. In general C
o
eq
< C
eq
.
_____________________________________________________________________________
_
EXAMPLE 22.4 Calcite growth kinetics
In the growth of calcite crystals from supersaturated solutions, it may be assumed that the "adsorbed lattice
ions combine in the surface layer to form an ion-pair, CaCO
3
o
, that is subsequently incorporated into the surface
lattice:
Ca
2+
+ CO
3
2-
= CaCO
3
o
(1)
The equilibrium constant corresponding to the above reaction is given by:
K
1
= {CaCO
3
o
}/{Ca
2+
}{CO
3
2-
} (2)
(a) Show that if the growth rate is first order with respect to the growth unit,, then the corresponding rate
law is:
-d[Ca
2+
]/dt = (kK
1
A
2
2
)([Ca
2+
][CO
3
2-
] - K
so
/
2
2
/ (3)
where k is the first order rate constant, A is the "area of seed per unit volume of solution",
2
2
is the
divalent ion activity coefficient, K
1
is the formation constant of CaCO
3
o
, and K
so
is the solubility
product of calcite.
(b) Verify that the data below are consistent with the above growth model
Solution
(a) Let us recall Equation 22.44:
J = J
d
= J
r
= k(C - C
eq
) (22.44)
For calcite crystal growth, this equation can be re-expressed as:
J = J
d
= J
r
= k([CaCO
3
o
] - CaCO
3
o
]
eq
) (4)
According to Equation 2,
{CaCO
3
o
} = K
1
{Ca
2+
}{CO
3
2-
} = K
1

2
2
[Ca
2+
][CO
3
2-
] (2)
If it assumed that the neutral species, CaCO
3
o
, has unit acivity coefficient, then it follows that
[CaCO
3
o
] = K
1
{Ca
2+
}{CO
3
2-
} = K
1

2
2
[Ca
2+
][CO
3
2-
] (5)
Also, at equilibrium,
[CaCO
3
o
]
eq
= K
1
{Ca
2+
}
eq
{CO
3
2-
}
eq
= K
1
K
so
(6)
Substitution of Equations 5 and 6 into Equation 4 gives the desired expression.

EXAMPLE 22.5 Silica growth kinetics
The kinetics of silica growth were investigated by Fleming (J. Colloid Interface Sci., 110, 40-64 (1986)).
Seeded growth experiments were conducted where monodisperse colloidal amorphous silica (Ludox) was
introduced into a supersaturated solution of silicic acid (Si(OH)
4
). Under the experimental conditions (25C, 0.1M
NaCl, pH 6-8), the equilibrium silica solubility (determined as silicic acid concentration) is 115 ppm.
It has been proposed that growth occurs via a condensation polymerization reaction between a silicic acid
molecule and a surface hydroxyl:
-SiOH(s) + Si(OH)
4
(aq) -Si-O-Si(OH)
3
(s) + H
2
O (1)
(a) Figure E22.5a presents a set of experimental data plotted as 1n (C-C
eq
) vs. time. Establish that this plot,
which assumes that C
eq
corresponds to the equilibrium solubility of amorphous silica (115 ppm), does not
conform to the mechanism described by Equation 1.
(b) Fleming argues that a distinction should be made between a silica surface coated with chemisorbed silicic
acid (i.e., -Si(OH)
3
) and the surface of stable amorphous silica. Thus the appropriate C
eq
value to use in
analyzing the experimental data is not C
o
eq
= 115 ppm (the value used in plotting Figure E22.5a), but rather
C
x
, the silicic acid concentration in equilibrium with chemisorbed silicic acid. Replot the Figure 1 data as
-dC/dt vs C and show that the rate is relatively low when silicic acid concentration is below ~200 ppm.
Above this concentration, however, the rate is linearly dependent on the silicic acid concentration.
(c) The transitional concentration of ~200 ppm represents the silicic acid concentration that is in equilibrium
with a silica surface coated with chemisorbed silicic acid. Show that replotting the experimental data with
C
eq
= 200 ppm leads to a trend that is consistent with the proposed growth mechanism (Equation 1).
Solution
(a) Let C represent silicic acid concentration. Then the rate law corresponding to the proposed growth
mechanism (Equation 1) is:
-dC/dt = k
1
S[-SiOH]C - k
-1
S (2)
where S is the surface area of silica per volume of solution, and [-SiOH] is the surface concentration of
hydroxyl groups.
When the solution is in equilibrium with the chemisorbed surface, dC/dt = 0 and C = C
eq.
. Thus,
k
-1
= k
1
S[-SiOH]C
eq
(3)
Accordingly Equation 2 can be re-written as:
-dC/dt = k
1
S[-SiOH](C - C
eq
) = k' (C - C
eq
) (4)
where k' = k
1
S [-SiOH]. Equation 4 may be integrated with the initial condition C = Cwhen t = 0, to give
the following integrated rate law:
ln (C - C
eq
) = -kt (5)
where k = k'C
o
sat
and C
o
sat
>> C
eq
.
In view of Equation 5, the fact that the rate data plotted in Figure E22.5a do not give straight lines means
that either the proposed mechanism is inappropriate or the pertinent C
eq
is not the solubility of stable
amorphous silica.
(b) Figure E22.5b shows the required plot.
(c) Figure E22.5c shows the required plot.
Figure E22.5 Silica precipitation kinetic data.
l
n

(
C
-
C
e
)

(
p
p
m
)
700
pH = 8.01
7.31
6.33
0 100 200 300 400
Time, minutes
500 600
3.6
3.8
4.0
4.2
4.4
4.6
4.8
5.0
5.2
5.4
5.6
A
s
= 4195 t 60 cm
2
/cm
3
I = 0.1 M NaCl
(a)
S
i
l
i
c
i
c

A
c
i
d

D
e
p
o
s
i
t
i
o
n

R
a
t
e
,

-

C
/

I

(
p
p
m
/
m
i
n
)
pH = 8.01
7.31
6.33
100 200 300
1
2
3
4
5
Region 2 Region 1
C
x
400
C
e
Silicic Acid Concentration, C
(ppm)
Increasing
Time
(b)
pH = 8.01
7.31
6.33
3
2
4
5
0 100
200 300 400
Time, minutes
l
n

(
C
-
C
X
)

(
p
p
m
)
(c)
Figure E22.5 (contd) Silica precipitation kinetic data.
According to Equations 22.44, the crystal growth rate is proportional to the concentration
difference (C - C
eq
). However growth rates of inorganic salts in aqueous solution frequently
follow the form J ~ (C - C
eq
)
x
where x = 1.5 - 2. Thus the actual situation is much more
complicated than the simple two-step process presented above. In fact, several physicochemical
processes occur simultaneously, e.g. dehydration of ions, counter-diffusion of rejected water
molecules, surface diffusion, surface nucleation, etc.
The reaction-controlled growth models may be divided into (a) mononuclear layer, (b)
polynuclear layer, and (c) surface dislocation models. The mononuclear layer mechanism views
crystal growth in terms of the build-up of one monolayer of the crystal lattice, followed by the
formation of another layer on top of the previous one, etc. A new layer can grow only when a
two-dimensional nucleus is formed. The energetics of this two-dimensional nucleus can be
developed by following an approach similar to that used above for two-dimensional (i.e.
homogeneous) nucleation. That is, recalling Equation 22.29, the free energy of two-dimensional
nucleation consists of two terms:
G = VG

+ A (22.46)
where V is the volume of the nucleus and A is its total surface area. If a disc-like nucleus is
considered, with radius r and thickness h, then Equation 22.46 becomes:
G = 2rh + r
2
hG

(22.47)
It can be shown that the critical radius (r
c
) is given by
r
c
= -/G

(22.48)
and that the corresponding critical energy barrier G
crit
is given by
G
crit
= hr
c
(22.49)
Comparing Equations 22.32 and 22.48, it can be seen that the radius of a critical two-
dimensional radius is half that of a critical three-dimensional radius. It follows from Equations
22.26 and 22.49 that:
G
crit
= h
2

m
/RT ln S (22.50)
Therefore in analogy with the case of homogeneous nucleation (Equation 22.37), the rate
of two-dimensional nucleation may be written as:
R
n
= k
o
exp (-h
2

m
/R
2
T
2
ln S) (22.51)
Comparing Equations 22.36 and 22.50, it can be seen that
dim) 2 ( G
dim) 3 ( G
crit
crit

= (22.52)
___________________________________________________________________________
EXAMPLE 22.6 Comparison of energy barriers for three- and two-dimensional nucleation.
For (a) S = 1.1, and (b) S = 1.2, compare the energy barriers for three- and two-dimensional nucleation for
an inorganic salt for which = 10
-1
Jm
-2
,
m
= 2x10
-29
m
3
mol
-1
, h = 5x10
-10
m; also kT = 4x10
-21
J.
Solution
G
crit
(3 dim)/G
crit
(2 dim) = 50/1 for S = 1.1
= 1.2/1 for S = 1.2
These results indicate that at low supersaturations the critical energy barrier for three-dimensional
nucleation is significantly higher than that for two-dimensional nucleation. In other words low supersaturations will
favor growth rather than nucleation.
____________________________________________________________________________________________
22.3.3 Polynuclear Growth
22.3.4 Spiral Growth
It is frequently observed that for a given growth rate, the surface nucleus model of crystal
growth requires higher supersaturations than are actually observed experimentally. This
observation suggests that there may be more growth sites than indicated by the layer-by-layer
mechanism. The surface dislocation growth model explicitly takes account of the fact that
imperfections exist on crystal surfaces and that these crystallographic features can significantly
influence interfacial reactions.
Figure 22.5 presents a schematic illustration of the surface of a growing crystal. The
growth of a crystal involves several steps which can be summarized as follows:
(a) Transport of ions from the bulk solution to the solution immediately adjacent to
the crystal surface.
(b) Transfer of ions from the solution to an adsorption layer.
(c) Transfer of ions from the solution or adsorption layer to a growth step.
(d) Transfer of ions from the solution, adsorption layer or growth step to a growth
site, i.e., a lattice position at a kink.
Figure 22.5 Schematic illustration of the surface of a growing crystal: (a) Smooth surface,
(b) edge or step, (c) kink site, (d) two dimensional nucleus.
Figure 22.6 Schematic illustration of a growth spiral showing the kink distance (x
o
), the
step distance (y
o
), and the step height (a) (After Nielsen).
If the rate of crystal growth is determined by step (d), then in order to obtain the rate
equation it is necessary to have quantitative information on the number density of growth sites,
i.e., the number of kink sites per unit area. According to the Burton, Cabrera and Frank (BCF)
theory of crystal growth, the sites needed for growth are provided by screw dislocations that
terminate in the surface of the crystal. The screw dislocations are in the form of a spiral
staircase, as illustrated in Figure 22.6. The main characteristics of the surface structure needed
for the crystal growth model are the kink distance (x
o
), the step distance (y
o
), and the step height
(a). It follows from Figure 22.6 that the density of kinks (n) in the crystal surface (i.e., the
number of kinks per unit area of crystal surface) is given by
n = (x
o
y
o
)
-1
(22.53)
The BCF theory gives the following expressions for x
o
and y
o
:
x
o
= a S
-1/2
exp (a
2
/kT) (22.54)
y
o
= 19a
3
/(kTlnS) (22.55)
where a is the step height, S is the saturation ratio (S = C/C
o
, where C and Co are respectively
the concentrations of the electrolyte in the supersaturated and the equilibrium saturated
solutions), and is the interfacial tension of the crystal/aqueous interface. Thus it follows from
Equation 22.53-22.55 that the density of kinks may be expressed as:
n = (22.56)
(22.57)
Equation 22.57 uses the approximation:
S
1/2
lnS (S-1) (22.58)
Consider the crystal growth of a symmetrical electrolyte MA. It shall be assumed that
step (d) (in particular, transfer of ions from the adsorption layer to the growth site) is rate-
determining. Thus the electrolyte in the solution is in equilibrium with the ions in the adsorption
layer, i.e.,
M
z+
(aq) = M
z+
(ad) (22.59)
K
+
=
+
/C
M
(22.60)
A
z-
(aq) = A
z-
(ad) (22.61)
K
-
=
-
/C
A
(22.62)
where
+
and
-
are the adsorption densities of the cation and anion, and K
+
and K
-
are the
respective adsorption constants.
Let us focus on the transfer of the growth units from the adsorption layer into a kink site:
M
z+
(ad) M
z+
(kink) (22.63)
A
z-
(ad) A
z-
(kink) (22.64)
k
+
k
+
k
-
k
-
It should be noted that M
z+
will deposit onto an A
z-
ion in the lattice. Similarly A
z-
will deposit
onto an M
z+
ion in the lattice. Let
+
be the fraction of the kink site with exposed M
z+
ions and
let
+
be the fraction of the kink site with exposed A
z-
ions. Then

+
+
-
= 1 (22.65)
It follows therefore that the rate equations for the deposition of M
z+
and A
z-
at the kink
sites are given by Equations 22.66 and 22.67 respectively:
r
+
= k
+

+

-
- k
+
(22.66)
r
-
= k
-

-

+
- k
-
(22.67)
where r
+
and r
-
respectively, the rate of deposition per kink for the cation and anion.
The formation of a stoichiometric MA crystal requires that cation and anion are
incorporated into the crystal structure at equal rates, i.e., r
+
= r
-
= r. Therefore, it follows from
Equations 22.65-22.67 that:

+
= (22.68)
and
r = r
+
= r
-
=' (22.69)
At equilibrium, r = r
+
= r
-
= 0. Accordingly Equations 22.66-22.67 respectively give
Equations 22.70-22.71:
k
+
' = k
+

o+

o-
/
o+
(22.70)
k
-
' = k
-

o-

o+
/
o-
(22.71)
where the subscript ("o") refers to equilibrium conditions. Similarly under equilibrium
conditions, Equation 22.69 gives
k
+
' k
-
'/k
+
k
-
=
o+

o-
(22.72)
Combination of Equations 22.69 and 22.72 gives:
r =
+ + +
+ + +
+ +

k k k
) ( k k
o o
(22.73)
If the adsorbed layer is approximately electroneutral, then
+

-
. Therefore Equation
22.73 becomes:
r =
( )
+ + +
+ +
+ + +

k k / ) k k (
2
o
2
(22.74)
Furthermore, if at equilibrium the fraction of a kink site occupied by the cation equals that
occupied by the anion (i.e.,
o+
=
o-
), then recalling the assumed electroneutrality of the
adsorbed layer (i.e.,
o+
=
o-
), Equations 22.70 and 22.71 may be combined to give

o+
= (k
+
' + k
-
')/(k
+
+ k
-
) (22.75)
Combination of Equations 22.74 and 22.75 gives:
r = (22.76)

k
+
k

k
+
+ k

( )
(
+
+
o+
)
+

o+
( )

+
+
o+
( )



1
]
1

k
+
k

k
+
+ k

( )



_
,


o+
( )
(22.77)
According to Equation 22.60,

+
= K
+
C
M
(22.78)

o+
= K
+
C
MO
(22.79)
Using Equations 22.78 and 22.79 in Equation 22.77 gives:
r = (22.80)
= (22.81)
where S is the saturation ratio (= C
M
/C
MO
)
Recalling that the density of kinks is given by Equation 22.57, the crystallization rate per
unit area will be given by

- d C
M
/ d t = n r =


( k
+
+ k
-
) 1 9 a
4
( / k T ) e x p ( a
2
/ k T
)
k
+
k
-
K
+
C
M O
1
1
]
( S - 1 )
2
( 2 2 . 8 2 )

EXAMPLE 22.7 Parabolic crystal growth rate
Liu and Nancollas (J. Cryst. Growth, 6, 281-289(1970)) investigated the kinetics of crystal growth of
calcium sulfate dihydrate. The following represents a selection of their results:
t(min) 0 5 15 30 40 50 60
[Ca
2+
]x10
2
(mol/L) 3.90 3.66 3.45 3.23 3.11 2.98 2.94
t(min) 80 100 120 140 160 180
[Ca
2+
]x10
2
(mol/L) 2 .81 2 .70 2 .62 2 .53 2 .50 2 .45
Show that these results obey the rate law:
-d[Ca
2+
]/dt = k([Ca
2+
] - [Ca
2+
]
eq
)
2
(1)
Solution
Equation 1 may be integrated with the initial conditions [Ca
2+
] = [Ca
2+
]
o
when t = 0, to give
1/([Ca
2+
] - [Ca
2+
]
eq
) - 1/([Ca
2+
]
o
- [Ca
2+
]
eq
) = kt (2)
Figure E22.7a shows a plot of the original data and Figure E22.7b shows a plot of these data according to Equation
2. The straight lines observed in Figure E22.7b confirm that the crystal growth of calcium sulfate dihydrate follows
the parabolic rate law given by Equation 1.
40
Time, minutes
4.0
3.0
2.0
80 120 160
C
o
n
c
e
n
t
r
a
t
i
o
n

x

1
0
2
40
Time, minutes
240
160
80 120 160
(
m


m
o
)
-
1


(
m
i


m
o
)
-
1

x

1
0
80
(a) (b)
Figure E22.7
22.4 Aggregative Growth
The evidence for aggregative growth includes the following: (a) TEM micrographs that
indicate that primary particles in the 10 to 1000 nm size range consist of aggregates of smaller
subparticles (e.g. TiO
2

by Santacesaria et al., gold by Uyeda et al., iron hydroxides by
Murphy et
al, van der Woude and deBruyn, Murphy et. al, zirconia by Bleier and Cannon, titania by
Dirksen and Ring, silica by Bogush and Zukoski), (b) observation that particle densities decrease
with increase in size (e.g. iron (III) hydroxide by von Gunten and Schneider), and (c) aggregative
growth models that are reasonably consistent with experimental data (e.g. silica from alkoxides,
Philipse; Bogush and Zukoski).
___________________________________________________________________________
EXAMPLE 22.8 Evidence for aggregative growth: Dissolution of colloidal Fe(OH)
3
It has been suggested by von Gunten and Schneider (J. Colloid Interface Sci., 145, 127-139 (1991)) that the
aggregative nature of colloidal precipitates can be probed by conducting carefully designed dissolution experiments.
The overall reaction describing the acid decomposition of colloidal Fe(OH)
3
can be expressed as:
(Fe(OH)
3
)
p
+ 3pH
+
= pFe
3+
+ 3pH
2
O (1)
The overall reaction consists of two main steps:
I. Disintegration (i.e. disaggregation):
(Fe(OH)
3
)
p
= (p/m)(Fe(OH)
3
)
m
(2)
II. Dissolution of primary particles:
(Fe(OH)
3
)
m
+ 3mH
+
= mFe
3+
+ 3m H
2
O (3)
20
30
40
50
60
120
10/HCl
t
1
/
2

[
s
e
c
]
Figure 22.8a
In synthesis experiments conducted by von Gunten and Schneider, Fe(II) solution was oxygenated in the
presence of tris (i.e., tris (hydroxymethyl)aminomethane). The size of the colloidal Fe(III) hydroxide particles
ranged from 10 in 0.5M tris to 140 in 0.1M tris.
Examine each of the following observations for its compatibility with an aggregative growth model.
(a) In a dissolution experiment conducted in 2M HCl, the half-life was found to be
independent of the hydrodynamic radii of the starting particles.
(b) Separate dissolution experiments were conducted with 10 and 120 particles. With both small and large
particles, the half-life decreased with increase in HCl concentration. However, while t
1/2
was independent
of particle size for 2M HCl, as the acid concentration was decreased from this level, t
1/2
(120) increased
relative to t
1/2
(10).
(c) Both acid-and ligand-promoted dissolution occur when HCl is used as the lixiviant whereas with HClO
4
,
only proton-promoted dissolution is significant. Thus with 120 particles, the half-life is respectively 1
min and 12 min for HCl and HClO
4
. A solution containing 120 particles was mixed with an equal
volume of 2M HClO
4
. One minute after the addition of HClO
4
, the pH was raised to 3 by addition of 3M
tris (pH 8). Following this HCl was added and the dissolution rate was monitored. The results are
presented in Figure E22.8a, where the 10/HCl data are also shown for comparison.
Solution
(a) The reported observation is illustrated schematically in Figure E22.8b
1.0 1.5 2.0
0
10
120/HClO4/HCl
HCl Concentration, M
r
[Tris], mol/L
t
1/2
0.1 0.5
Figure E22.8b
Since dissolution is a heterogeneous process, the rate is expected to be proportional to the external surface
area. The fact that the rate was found to be independent of the "external" surface area (as determined by
the hydrodynamic radius) indicates that this is not the relevant external surface area. This result implies
that if (i) the primary particles are aggregates of equal-size subparticles, and (ii) the disintegration step
(Equation 2) is extremely fast, then the external surface seen by the aqueous solution corresponds to that
provided by the subparticles. Therefore the half-life should be constant for all sizes of the primary
particles, since in all cases the actual dissolving particles are the subparticles (i.e.,(Fe(OH)
3
)
m
, see
Equation 3).
(b) A schematic illustration of this observation is shown in Figure E22.8c. Two effects are at work here. With
2M HCl the acid concentration is high enough to permit fast disaggregation of the 120 particles. Thus in
this case the hydrodynamic radius of the starting particles will have no effect on the half-life. However, as
the acid concentration is decreased the disaggregation step becomes increasingly slow and the external
surface area increasingly has contributions from the outer surface delimited by the hydrodynamic radius.
(c) During the preliminary HClO
4
treatment, the primary particles disaggregate to give the subparticles, but
because of the slow dissolution kinetics with this acid (compared with HCl), very little actual dissolution
(i.e., step II, Equation 3) occurs. In the subsequent reaction in HCl, the subparticles generated in the
pretreatment undergo dissolution. The fact that the dissolution behavior coincides with that of the
10/HCl system strongly suggests that the subparticles (i.e. (Fe(OH)
3
)
m
) are of the order of 10 in radius.
[HCl]
t
1/2
2 M
10A
o
120 A
o
Figure E22.8c
____________________________________________________________________________________________
The sequence of steps involved in aggregative growth are illustrated in Figure 22.7.
Quantitative analysis of the overall process, taking into account all the separate steps, is a
formidable task. However, by means of carefully designed seeded growth experiments, it is
possible to obtain useful quantitative insight into the factors that control an aggregative growth
process.
Consider an experimental system where a supersaturated solution containing initially
contains a number concentration n
o
of subparticles (nuclei). At t = 0, a known amount of
primary particles is introduced into the solution. The primary particles are much larger than the
subparticles. Let r
1
= radius of a primary particle, r
2
= radius of a subparticle. It is assumed that
(a) the primary particles do not aggregate (e.g. presence of significant electrostatic repulsion),
(b) no nucleation occurs during the seeded growth process, (c) particle growth may be viewed in
terms of the coagulation of small subparticles on a large stationary central primary particle.
Figure 22.7 Schematic illustration of the stages of aggregative growth (after Bagnall et.al.
1990).
Let us consider a spherical surface located at a distance r from a central primary particle.
Figure 22.8 illustrates this situation. The flux of subparticles towards the central primary
particle is given by
J = DC/r (22.83)
The number of subparticles crossing this shell in a given time is given by
R = 4r
2
J (22.84)
At steady-state, the number of subparticles crossing the shell in a given time, must equal the
number of subparticles colliding with the central sphere in the same time interval, i.e.,
R = (dn/dt)
r=r
1
= (dn/dt)
r
= 4r
2
J = constant (22.85)
r
r
2
r
12
r
1
SUBPARTICLE
PRIMARY
PARTICLE
Figure 22.8 Schematic illustration of the subparticle (r
2
)-primary particle (r
1
)
aggregation (after Philipse, 1988)
Recalling Equation 9.47, and the assumption that r
12
= r
1
+ r
2
r
1
, the flux may be
expressed as
J = Dr
12
n

/r
2
Dr
1
n

/r
2
(22.86)
where n

is the concentration of subparticles at infinite distance from the central primary


particle. Substituting Equation 22.86 into 22.85,
R = (4r
2
)(Dr
1
n

/r
2
) = 4Dn

r
1
(22.87)
Since the central absorbing sphere is also subject to Brownian motion, the appropriate
expression for the diffusion coefficient is
D = D
1
+ D
2
(22.88)
where D
1
and D
2
respectively represent the diffusion coefficients of a primary particle and a
subparticle. However, since it is assumed that the primary particles are much larger than the
subparticles, the diffusion coefficient of the primary particles may be reasonably taken to be
relatively small, i.e.,
D D
2
(22.89)
The central sphere grows by absorbing the subparticles. Let V
1
and V
2
respectively be
the volumes of a primary particle and a subparticle. Then it follows that
dV
1
/dt = V
2
R (22.90)
Recalling that V
1
= 4r
1
3
/3, Equation 22.90 can be rewritten as:
4r
1
2
dr
1
/dt = V
2
R (22.91)
Combining Equations 22.87 and 22.91 gives:
4r
1
2
dr
1
/dt = 4Dn

r
1
V
2
(22.92)
That is,
r
1
dr
1
= Dn

V
2
dt (22.93)
Infinite Volume of Aqueous Phase. If the assumption is made that the aqueous phase is of
infinite volume, then the bulk concentration of subparticles is constant. Then Equation 22.93
may be integrated with the initial condition, r
1
= r
o
at t = 0 to give:
r
1
2
= r
o
2
+ 2V
2
n

D
2
t (22.94)
Finite Volume of Aqueous Phase. When the aqueous phase has a finite volume, the
possibility exists that the subparticles may become exhausted in the course of the growth
process. Let n
po
represent the initial number concentration of primary particles, each with an
initial radius of r
o
. At a given time, t, when each primary particle has attained a size of r
1
, the
prevailing bulk concentration of subparticles (n

) will be
n


= n
o
- (4/3)n
po
(r
1
3
- r
o
3
)/V
2
(22.95)
where n
o
is the initial number concentration of subparticles.
The particles attain their maximum size (r
m
) when the bulk concentration of subparticles
becomes exhausted. Therefore inserting n

= 0 in Equation 22.95 gives:


n
o
= (4/3)n
po
(r
m
3
- r
o
3
)/V
2
(22.96)
Combining Equations 22.95 and 22.96,
n

= (4/3)n
po
(r
m
3
- r
1
3
)/V
2
(22.97)
The growth of each sphere may be described with Equation 22.93, where n

is now
given by Equation 22.97. It can be shown (see Philipse, 1988) that Equations 22.93 and 22.97
give the following result:
F(x) = F(x
o
) + 8r
m
n
po
D
2
t (22.98)
where
x = r
1
/r
m
, x
o
= r
o
/r
m
(22.99)
and
F(x) 1n
1+ x + x
2
x 1 ( )
2



1
]
1
2 3 tan
1
2x +1
3


_
,
(22.100)
___________________________________________________________________________
EXAMPLE 22.9 Aggregative growth: Infinite vs finite volume of aqueous phase.
In a seeded growth experiment, silica particles grow from an initial radius of 60 nm to 80 nm in 1-2h. The
volume fraction of subparticles is initially 0.005. The diffusion coefficient for a subparticle is of the order of 10
-5
cm
2
s
-1
. Determine whether the observed particle growth rate is consistent with an infinite volume of aqueous
phase.
Solution
For an infinite volume of aqueous phase, the aggregative growth obeys Equation 22.94:
r
1
2
= r
o
2
+ 2V
2
n

D
2
t (22.94)
That is,
t = (r
1
2
- r
o
2
)/2V
2
n

D
2
(1)
In this expression the term V
2
n

represents the volume fraction of subparticles, i.e., V


2
n

= 0.005; also r
1
= 80nm,
r
o
= 60nm, D
2
= 10
-5
cm
2
s
-1
. Thus
t = [(80 nm)
2
- (60 nm)
2
]/2 (0.005)(10
-5
cm
2
s
-1
) = 5.1 x 10
-4
s
The observed time was 1-2h, which is much greater than the theoretical value of 5.1x10
-4
s based on an
infinite volume of aqueous phase. Thus it may be concluded that the silica growth did not occur in an infinite
volume of aqueous phase.
____________________________________________________________________________________________
22.5 Synthesis of Monodispersed Colloidal Particles
In the monomer-addition growth model, the formation of monodispersed colloidal
particles is rationalized with the aid of the LaMer diagram, as shown in Figure 22.9. Three
stages are identified. Stage I involves slow monomer generation (e.g. via decomposition of
precursor compounds). No particle formation occurs during this stage. Stage II represents the
nucleation stage. Nucleation commences when the monomer concentration exceeds C. At first
the monomer concentration continues to rise since monomers are produced more rapidly than
they are consumed by nucleation. However, eventually there is a depletion of monomers as
monomer generation is no longer able to keep up with its consumption by nucleation. When C
falls below C
*
min
, no more nucleation occurs. Stage III represents the growth stage; here the
predominant process is the addition of monomers to the available nuclei.
I
II III
C*
max
C*
min
C
s
CONC'N (C)
TIME (t)
Figure 22.9 LaMer diagram illustrating the stages in the formation of monodispersed
colloidal particles.
C
*
max
represents the critical limiting supersaturation. In order to obtain monodisperse
particles, the time lapse between C
*
min
and C
*
max
must be as small as possible. That is, in order
to obtain monodispersed particles, there must be a clear temporal separation between the
nucleation and growth stages.
EXAMPLE 22.15 Critical supersaturation during homogeneous precipitation of metal sulfides with
thioacetamide
The precipitation of a divalent metal sulfide (MS) can be expressed as:
M
2+
+ S
2-
= MS (s) (1)
The sulfide ions needed for the precipitation reaction may be generated in-situ via the thermal decomposition of
thioacetamide (TA):
CH
3
C(S)NH
2
+ H
2
O CH
3
C(O)NH
2
+ H
2
S (2)
The corresponding rate law is given by,
-d[TA]/dt = k[H
+
] [TA] (3)
Taking into consideration the deprotonation reactions of H
2
S,
H
2
S = HS
-
+ H
+
K1 (4)
HS
-
= S
2-
+ H
+ K
2 (5)
show that prior to precipitation, the time dependence of sulfide concentration is given by:
[S
2-
] = [TA]
o
{1 - exp(-k[H
+
]t)}/{[H
+
]
2
/K
1
K
2
+ [H
+
]/K
2
+ 1} (6)
Solution
It follows from the rate law (Equation 3) that,
[TA]
o
- [TA]
t
= [TA]
o
{1 - exp(-k[H
+
]t)} (7)
Also, it must be recognized that, according to the stoichiometric relationships in Equation 2, the total dissolved
sulfide corresponds to the amount of decomposed thioacetamide. Therefore,
[TA]
o
- [TA]
t
= [S]
T
(8a)
= [H
2
S] + [HS
-
] + [S
2-
] (8b)
Combining Equations 4 and 5, we get:
H
2
S = S
2-
+ 2H
+
K
1
K
2
(9)
Thus,
[H
2
S] = ([H
+
]
2
/K
1
K
2
)[S
2-
] (10)
Also, from Equation 5,
[HS
-
] = ([H
+
]/K
2
) [S
2-
] (11)
It follows from Equations 10 and 11 that
[S]
T
= [H
2
S] + [HS
-
] + [S
2-
]
=( [H
+
]
2
/K
1
K
2
+ [H
+
]/K
2
+ 1 ) [S
2-
] (12)
That is,
[S
2-
] = [S
T
]/([H
+
]
2
/K
1
K
2
+ [H
+
]/K
2
+ 1) (13)
It follows from Equations 7, 8a and 13 that
[S
2-
] = [TA]
o
{1-exp(-k[H
+
]t}/{[H
+
]
2
/K
1
K
2
+ [H
+
]/K
2
+ 1}
(See Celikkaya and Akinc, JACerS, 73, 2360 (1990))

FURTHER READING
References on Crystallization and Precipitation
1. J. Nyvlt, O. Sohnel, M. Matuchova, and M. Broul, The Kinetics of Industrial
Crystallization, Elsevier, New York, NY, 1985.
2. J. W. Mullin, Crystallization, 2nd Ed., CRC Press, Cleveland, OH, 1972.
3. W. L. McCabe and J. C. Smith, Unit Operations of Chemical Engineering, 3rd Ed.,
McGraw-Hill, New York, 1976.
4. L. Gordon, M. L. Salutsky and H. H. Willard, Precipitation from Homogeneous Solution,
Wiley, New York, 1959.
5. G. H. Nancollas, "The Growth of Crystals in Solution", Adv. Colloid Interface Sci., 10,
215-252(1979).
6. G. H. Nancollas and N. Purdue, "The Kinetics of Crystal Growth", Quart. Rev., 18, 1-
20(1964).
7. A. E. Nielsen, "Theory of Electrolyte Crystal Growth. The Parabolic Rate Law", Pure
Appl. Chem., 53, 2025-2039(1981).
8. A. E. Nielsen, "Electrolyte Crystal Growth Mechanisms", J. Cryst-Growth, 67, 289-
310(1984).
9. A. S. Myerson, ed., Crystallization as a Separations Process, ACS Symp. Ser., Vol. 438,
1990.
10. A. E. Nielsen, Kinetics of Precipitation, Macmillan, New York, 1964.
11. A. G. Walton, The Formation and Properties of Precipitates, Interscience, New York,
1967.
12. P. Hartman, ed., Crystal Growth: An Introduction, North-Holland, Amsterdam, 1973.
13. R. F. Strickland-Constable, Kinetics and Mechanism of Crystallization, Academic, London,
1968.
References on Hydrogen Reduction
1. F. A. Schaufelberger and T. K. Roy, "Separation of Copper, Nickel and Cobalt by
Selective Reduction from Aqueous Solution", Trans. IMM, 64, 375-93 (1955).
2. F. A. Schaufelberger, "Precipitation of Metal from Salt Solution by Reduction with
Hydrogen", Trans. AIME, , 539-548 (1956).
3. B. Meddings and V. N. Mackiw, "The Gaseous Reduction of Metals from Aqueous
Solution", in Unit Processes in Hydrometallurgy, M. E. Wadsworth and F. T. Davis, eds.,
Gordon and Breach, New York, NY, 1964, pp. 345-384.
4. A. R. Burkin, "Production of Metal Powders and Coatings by Precipitation Techniques,
and their Fabrication", Met. Revs.,12, 1-14 (1967).
5. D. J. I. Evans, "Production of Metals by Gaseous Reduction from Solution - Processes
and Chemistry", in Advances in Extractive Metallurgy, IMM, London, 1968, paper 35.
6. J. Halpern, "Homogeneous Catalytic Activation of Molecular Hydrogen by Metal Ions
and Complexes", J. Phys. Chem., 63, 398-403 (1959).
7. M. Findlay, "The Use of Hydrogen to Recover Precious Metals", in Precious Metals
1982, M. I. El Guindy, ed., Pergamon, New York, NY, 1983, pp. 477-501.
References on Sulfide Precipitation
1. C. S. Simons, "Hydrogen Sulfide as a Hydrometallurgical Reagent", in Unit Processes in
Hydrometallurgy, M. E. Wadsworth and F. T. Davis, eds., Gordon and Breach, New
York, NY,1964, pp. 592-616.
2. T. K. Roy, "Preparing Nickel and Cobalt Concentrates" Ind and Eng Chem., 53, 559-566
(1961).
3. S. Joris, "La Cintique de prcipitation des Sulfures de Cobalt et de Nickel par
L'Hydrogne Sulfur," Bull. Soc. Chim. Belges., 78, 607-619 (1969).
4. H. A. Pohl, J. Am. Chem. Soc., 76, 2182-2184 (1954).
5. M. C. Jha and G. A. Meyer, "Physical Chemistry of Nickel Sulfide Precipitation from
Acidic Sulfate Solutions", TMS Paper Selection, A80-51, 1980.
6. M. C. Jha, G. A. Meyer, and G. R. Wicker, "An Improved Process for Precipitating
Nickel Sulfide from Acidic Laterite Leach Liquors", J. Metals, November, pp. 48-53,
1981.
7. B. Tougarinoff, "Dnickelage du Cobalt par Cmentation Sulfurante," Ind. Chim. Belge,
Tome XX, Numro Spcial, Vol. II, pp. 532-536 (1955).
References on Electroless Plating
1. F. Pearlstein, "Electroless Plating", in Modern Electroplating, F. A. Lowenheim, ed., 3rd
Ed., Wiley, New York, 1974, pp. 710-746.
2. A. K. Graham, ed., Electroplating Engineering Handbook, 3rd Ed., Van Nostrand
Reinhold, New York, 1971, Chapt 15, "Nonelectrolytic Metal Coating Processes", pp.
475-507.
3. G. O. Mallory and J. B. Haydu, eds., Electroless Plating - Fundamentals and
Applications, American Electroplaters and Surface Finishers Society, Orlando, FL, 1990.
4. Y. Okinaka and T. Osaka, "Electroless Deposition Processes: Fundamentals and
Applications," in Advances in Electrochemical Science and Engineering, Vol. 3, pp. 55-
116 (1994).
References on Aggregative Growth
1. E. Santacesaria, M. Tonello, G. Storti, R. C. Pace, and S. Carra, "Kinetics of Titanium
Dioxide Precipitation by Thermal Hydrolysis", J. Colloid Interface Sci., 111, 44-53
(1986).
2. A. P. Philipse, "Quantitative Aspects of the Growth of (Charged) Silica Spheres", Prog.
Colloid Polym. Sci., 266, 1174-1180 (1988).
3. C. M. Bagnall, L. G. Howarth, and P. F. James, "Modelling of Aggregation Kinetics of
Colloidal Silica Particles", J. Non-cryst. Solids, 121, 56-60 (1990).
4. G. H. Bogush and C. F. Zukoski, "Uniform Silica Particle Precipitation: An Aggregative
Growth Model", J. Colloid Interface Sci., 142, 19-34 (1991).
5. J. Y. Bottero, M. Axelos, D. Tchoubar, J. M. Cases, J. J. Fripiat, and F. Fiessinger,
"Mechanism of Formation of Aluminum Trihydroxide from Keggin Al
13
Polymers", J.
Colloid Interface Sci., 117, 47-57 (1987).
6. A. Bleier and R. M. Cannon, "Nucleation and Growth of Uniform m-ZrO
2
", in Better
Ceramics Through Chemistry II, C. J. Brinker, D. E. Clark, and D. R. Ulrich, eds., MRS,
Pittsburgh, PA, 1986, pp. 71-78.
7. J. A. Dirksen and T. A. Ring, "Production of Powders for High-tech Ceramics", in High-
tech Ceramics: Viewpoints and Perspectives, G. Kostorz, ed., Academic, New York,
1989, pp. 29-39.
8. T. Sugimoto and E. Matijevic, "Formation of Uniform Spherical Magnetite Particles by
Crystallization from Ferrous Hydroxide Gels", J. Colloid Interface Sci., 74, 227-243
(1980).
9. W. P. Hsu, L. Ronnguist, and E. Matijevic, "Preparation and Properties of
Monodispersed Colloidal Particles of Lanthanide Compounds. 2. Cerium (IV)",
Langmuir, 4, 31-37 (1988).
10. J. H. A. van der Woude, J. B. Rijnbout, and P. L. de Bruyn, "Formation of Colloidal
Dispersions from Supersaturated Iron (III) Nitrate Solutions, IV. Analysis of Slow
Flocculation of Goethite", Colloids Surf., 11, 391 - 400 (1984).
11. P. J. Murphy, A. M. Posner, and J. P. Quirk, "Characterization of Partially Neutralized
Ferric Nitrate Solutions", J. Colloid Interface Sci., 56, 270-283 (1976); "Characterization
of Partially Neutrallized Ferric Chloride Solutions", J. Colloid Interface Sci., 56, 284-297
(1976); "Characterization of Partially Neutralized Ferric Perchlorate Solutions", J.
Colloid Interface Sci., 56, 298-311 (1976); "Characterization of Hydrolyzed Ferric Ion
Solutions. A Comparison of the Effects of Various Anions on the Solutions", J. Colloid
Interface Sci., 56, 312-320 (1976).
12. K. M. Towe and W. F. Bradley, "Mineralogical Constitution of Colloidal Hydrous Ferric
Oxides", J. Colloid Interface Sci., 24, 384-392 (1967).
13. N. Uyeda, M. Nishino, and E. Suito, "Nucleus Interaction and Fine Structures of
Colloidal Gold Particles", J. Colloid Interface Sci., 43, 264-276 (1973).
References on Synthesis of Monodispersed Particles
1. E. Matijevic, "Production of Monodispersed Colloidal Particles", Ann. Rev. Mater. Sci.,
15, 483-516 (1985).
2. T. Sugimoto, "Preparation of Monodispersed Colloidal Particles", Adv. Colloid Interface
Sci., 28, 65-108 (1987).
3. M. Haruta and B. Delmon, "Preparation of Homodisperse Solids", J. Chem. Phys., 83,
859-868 (1986).
4. J. Th. G. Overbeek, "Monodisperse Colloidal Systems, Fascinating and Useful", Adv.
Colloid Interface Sci., 15, 251-277 (1982).
5. T. Sugimoto, "Preparation and Characterization of Monodispersed Colloidal Particles",
MRS Bull., Dec., 23-28 (1989).
6. J. Livage, M. Henry, J. P. Jolivet, and C. Sanchez, "Chemical Synthesis of Fine
Powders", MRS Bull., Jan., 18-25 (1990).
7. M. Ozaki, "Preparation and Properties of Well-defined Magnetic Particles", MRS Bull.,
Dec., 35-40 (1989).
8. V. K. LaMer and R. H. Dinegar, "Theory, Production and Mechanism of Formation of
Monodispersed Hydrosols", J. Amer. Chem. Soc., 72, 4847-4854 (1950).
9. K. Osseo-Asare, Microemulsion-Mediated Synthesis of Nanosize Oxide Materials, in
Handbook of Microemulsion Science and Technology, P. Kumar and K. L. Mittal, eds., Marcel
Dekker, New York, NY, 1999, pp. 549-603.
10. K. Osseo-Asare and F. J. Arriagada, Growth Kinetics of Nanosize Silica in a Nonionic
Water-in-Oil Microemulsion: A Reverse Micellar Pseudophase Reaction Model, J. Colloid
Interface Sci., 218, 63-76 (1999).
11. T. Sugimoto, ed., Fine Particles. Synthesis, Characterization, and Mechanisms of
Growth, Marcel Dekker, New York, NY, 2000.
References on Electrodeposition
1. J. O'M. Bockris and G. Razumney, Fundamental Aspects of Electrocrystallization,
Plenum, New York, 1967.
2. A. Brenner, Electrodeposition of Alloys, Vols. I and II, Academic, New York, 1963.
3. M. Fleishmann and H. R. Thirsk, "Metal Deposition and Electrocrystallization", in P.
Delahay and C. Tobias, eds., Advances in Electrochemistry and Electrochemical
Engineering, Vol. 3, 1963, p. 123.
4. A. Damjanovic, "The Mechanism of the Electrodeposition of Metals", in J. O'M. Bockris
and B. E. Conway, eds., Modern Aspects of Electrochemistry, Vol. III, 1964, p. 224.
5. J. O'M. Bockris and A. K. N. Reddy, Modern Electrochemistry, Vol. 2, Plenum New
York, 1970, pp. 1173 - 1231.
6. L. Young, Anodic Oxide Films, Academic, New York, 1961.
7. A. Calusaru, Electrodeposition of Metal Powders, Elsevier, New York, 1979.
8. M. Paunovic and M. Schlesinger, Fundamentals of Electrochemical Deposition, Wiley,
New York, 1998.

Das könnte Ihnen auch gefallen