Preview-9780203908280 A23566891
Preview-9780203908280 A23566891
TECHNOLOGY
HANDBOOK
CRYSTAlllZATION
TECHNOlOGY
HANDBOOK
Second Edition
Revised and Expanded
edited by
A. Mersmann
Technische Universitiit Munchen
Garching, Germany
0 ~~~,~~!~ro,p
Boca Raton London New York
Headquarters
Marcel Dekker, Inc.
270 Madison Avenue, New York, NY 10016
tel: 212---696-9000; fax: 212-685-4540
The publisher offers discounts on this book when ordered in bulk quantities. For more infor
mation, write to Special Sales/Professional Marketing at the headquarters address above.
Neither this book nor any part may be reproduced or transmitted in any form or by any
means, electronic or mechanical, including photocopying, microfilming, and recording, or
by any information storage and retrieval system, without permission in writing from the
publisher.
Preface
The aim of this book is to provide reliable information not only on the
science of crystallization from solution and from melt but also on the
basic design methods for laboratory and especially industrial crystallizers.
Up to now the niche between scientific results and practical design and
operation of large-scale crystallizers has scarcely been filled. A work devoted
to this objective has to take into account relevant crystallization phenomena
as well as chemical engineering processes such as fluid dynamics, multiphase
flow, and heat and mass transfer. In the design of crystallizers, experiments
are initially performed on laboratory crystallizers to obtain kinetic data. In
this book, information is given on reliable scale-up of such crystallizers. The
selection, design, and operation of large-scale industrial crystallizers based
on fundamentals is the most significant objective of this work. To this end,
an appendix listing important physical properties of a large number of
crystallization systems is included. A selection of design data valid for indus
trial crystallizers with volumes up to several hundred cubic meters demon
strates the applicability of the design and scale-up rules.
iii
iv Preface
To date, the design of crystallizers has not been achievable from first
principles. The reason is very simple: a complex variety of different pro
cesses occur in crystallizers, such as nucleation, crystal growth, attrition and
agglomeration of crystals, fluid dynamics, and heat and mass transfer. Some
of these processes are not yet well understood although the design and
operation of large-scale crystallizers require reliable knowledge of the
most essential processes. The book presents a survey of the state of the
art and stresses the interrelationships of the essential mechanisms in such
an apparatus. Furthermore, with respect to nucleation and crystal growth,
general approaches have been developed to predict the kinetic rates that are
needed for chemical engineering design and a new chapter on agglomeration
has been added.
Supersaturation is the decisive driving force with respect to the kinetics of
crystallization. Optimal supersaturation is a prerequisite for the economical
production of crystals with a desired size, shape, and purity. The book offers
information on the most suitable supersaturation requirements in labora
tory and industrial crystallizers. Not only are aspects of cooling and
evaporative crystallization considered, but drowning-out and reaction crys
tallization are also described in detail. In dealing with precipitation the
complex interrelationships between mixing and product quality are dis
cussed. A special segment is devoted to the problem of how the process
components of an entire crystallization process can be economically fitted
together. The aspects stressed are always those of production quality; size
distribution, coefficient of variation, crystal shapes and purity, and the
problem of encrustation are considered. One chapter is devoted to the con
trol of crystallizers and another deals with the role of additives and impu
rities present in the solution. Crystallization from the melt is described in full
detail, and information is given on how to design and operate the corre
sponding crystallizers. The book describes the most significant devices for
crystallization from the melt and solidification processes. Process develop
ment such as high-pressure crystallization and freezing are considered and in
this second edition new results on direct contact cooling crystallization have
been added.
My goal has been to edit a work as homogeneous and practical as pos
sible. The book is therefore not a mere collection of independent articles
written by several contributors but a coordinated handbook with a single list
of symbols and a unified bibliography. It is divided into 15 chapters to make
it easier to find points of interest. Only simple derivations and equations
absolutely necessary for understanding and for calculation are presented.
The book is based on literature that is available worldwide (especially
references from the United States, Europe, and Japan) and on the direct
experience of the contributors. Some of the contributors work in the indus-
Preface V
trial sector, and nearly all have spent some time in industrial plants. A few
fundamental chapters were written by scientific researchers at universities.
Because this volume addresses the theory and practice of crystallization, it
should be valuable in both academia and industry.
A. Mersmann
Contents
Preface iii
Contributors xiii
vii
viii Contents
2. Activated Nucleation 45
A. Mersmann, C. Heyer, and A. Eble
1. Homogeneous Nucleation 47
2. Heterogeneous Nucleation 58
3. Surface Nucleation on Crystals 67
4. Comparison of Nucleation Rates 71
References 76
3. Crystal Growth 81
A. Mersmann, A. Eble, and C. Heyer
6. Agglomeration 235
A. Mersmann and B. Braun
6. Incrustation 437
7. Fitting the Process Parts Together 450
References 471
10. Challenges in and an Overview of the Control of Crystallizers 475
S. Rohani
1. Introduction 475
2. Dynamic Modeling of Crystallization Processes 479
3. Instrumentation in Crystallization Control 484
4. Control of Crystallization Processes 490
5. Conclusion 508
References 509
11. Reaction Crystallization 513
R. David and J. P. Klein
1. Introduction 513
2. Driving Force of Reaction Crystallization 515
3. Reaction Crystallization Kinetics 520
4. Fluid Dynamics, Mixing, and Precipitation 530
5. Conclusion: A General Methodology to Solve a Reaction
Crystallization Problem 555
References 557
12. "Tailor-Made Additives" and Impurities 563
I. Weissbuch, L. Leiserowitz, and M. Lahav
1. Introduction 563
2. Tailor-Made Additives for Crystal Morphology Engineering 564
3. Tailor-Made Additive Molecules for Crystal Dissolution;
Stereospecific Etchants 572
4. Theoretical Modeling 587
5. Mode of Occlusion of Impurities in Crystals 591
6. Tailor-Made Additives for Inhibition and Promotion of
Crystal Nucleation 602
7. Outlook 612
References 612
13. Suspension Crystallization from the Melt 617
K. Toyokura and I. Hirasawa
1. Introduction 617
Contents xi
Appendix 725
Notation 787
Bibliography 799
Substance Index 807
Subject Index 819
Contributors
Unlike crystallization from melting and freezing, heat transfer is not the
decisive process in cystallization from solution because the properties of
the crystalline product depend primarily on supersaturation generated by
cooling, evaporation, drowning out, or a reaction. The quality components
of the product (i.e., its crystal size distribution, its median crystal size, its
purity, and its crystal shape) are strongly influenced by (a) the geometry and
type of the crystallizer, (b) the operating conditions, and (c) the properties of
the liquid and solid phases.
The requirements of the product rather than the method of creating
supersaturation are decisive when selecting a crystallizer. In an industrial
crystallizer, there is such a variety of complex processes that it is difficult for
the chemical engineer to decide on a suitable procedure for the design of a
full-scale apparatus. Therefore, let us consider some general guidelines.
If the aim is to obtain a certain product having a specific crystal size
distribution, grain size, and purity instead of a random product, it is
necessary to control the local and mean supersaturation as well as the
residence time of the solid in the supersaturated solution. Supersaturation
1
2 Mersmann
is a prerequisite for nucleation and growth, which are decisive not only for
the formation of a solid phase but also for its occurrence (i.e., size distribu
tion of crystals and their shape). The degree of supersaturation is deter
mined by the flows of materials and energies, on the one hand, and by
crystallization kinetics, such as nucleation and growth, on the other hand.
In addition to the conservation laws of materials and energy, the population
balance is very important because in most cases, a crystalline product with a
certain size distribution is required. With respect to this fact, it is necessary
to take into account processes that influence the population balance (i.e.,
mainly agglomeration for very small crystals and attrition for large crystals).
As a rule, crystals have a higher density than that of the surrounding
liquid, which results in settling. Therefore, a certain upflow current is needed
in all crystallizers to compensate settling. As a result, crystallizers are
equipped with rotors, such as stirrers or pump impellers, which can cause
attrition of large crystals. In addition, the quality of a crystalline product
may depend on the fluid dynamics of the slurry in the crystallizer. As a rule,
the distribution of supersaturation and solid material in a crystallizer
depends on the process of macromixing. This may be important for coarse
crystalline products. With drowning-out and reaction crystallization, the
local supersaturation is influenced by the process of micromixing (i.e., mix
ing on a molecular scale). Finally, the production rate of an industrial
crystallizer may be high or low. As a rule, continuous_ crystallizers are
used in the case of high production rates because this operation is more
economical with respect to investment, energy, and labor costs. If several
products are to be crystallized in the same crystallizer, a batch crystallier is
chosen. Sometimes, crystallizers with circulating slurry or with fluidized
beds are used instead of stirred vessels. As will be shown later, the fluid
dynamics and, consequently, phenomena such as mixing and attrition differ
substantially in such apparatus. Therefore, it is necesary to provide detailed
information on the various crystallizers and their flow behavior. As stated
earlier, the most important process parameter is supersaturation, or the
difference between the actual concentration and the equilibrium concentra
tion of the liquid.
c - CM
- [kmol
m3 •
kg ] [kg
kmol or m 3 -
g g]
dm 3 - I (1.2)
k k
Total mass M =Ma+Mh M = LMi Total amount of n = na + nb n= Lni
)=a substance J=a
Ma M; X = na
Mass fraction U' =- W·=- Mole fraction Xj=~
" M I M n n
W =Ma n;
Mass ratio W-=_!!.J._ Mole ratio X =:.:E. X:-= -
a Mb 1 a nb 1
Mcarrier ncarrier
Conversion Mass ratio Mass fraction Mole ratio ~-----> Mole fraction
IV IV X
IV=-- w=--- X= X
w I- l+IV 1-x X= l+X
~ Af (IV l-w 0 ~ k ( k (
wj ]-1
Mean molar mass: Mm= - [
= --d!- +--_- )-I ; general: Mm= L [ ·y )]-I Mean density: Pm = L
m A1a Mb i=a Mi i=a PJ ~
~
;;i
3
~
=
=
Physical and Chemical Properties 5
the crystalline phase integrates (e.g., solvents, as is the case with all
hydrates). The ratios kg anhydrate/kg solvent and kg hydrate/kg solvent
are always different and, therefore, so are the supersaturation values. The
dimensionless supersaturations S and a also often vary considerably
depending on whether, for example, values of a 1 = flC/C*, a2 = fly/y*,
a 3 = fl Y / Y", a4 = fl w/w*, or a 5 = fl W / W* are involved. These statements
also show that the expression of supersaturation as a percentage is entirely
insufficient.
If the phase involved is a vapor or gaseous supersaturated phase, it is
often useful to use partial pressures instead of concentrations. If the ideal
law of gases applies, the following is valid:
S=.E_=f!_ (1.7)
C* Po
and
p-po
a =po
- (1.8)
Generally, it can be said that the difference in chemical potential flµ, = µ,F
-µ,c between the fluid (index F) and the cryal (index C) phases is the kinetic
driving force, which can be described via the relationship among the che
mical potential µ,, standard potential µ,0 , the activity a, and the heat of
crystallization flHcL:
µ, = Jl,o + fflTln a (l .9)
or
flµ, * = -flµ,
91T
= v In (a)
-
a*
= v ln Sa (solution) (l. lOa)
(1.11)
In this equation, z is the valency of the ions. The solubility product can be
formulated with activities a = y • y (y is the activity coefficient and Y; the
mole fraction of component i) or concentration C:
(1.12a)
or
(1.12b)
(1.13a)
(1.13b)
Solubility product
With activity Ka= (alr(ajJf = bAYAr(,BYBY
With concentration Kc= (c;.r(csY
C* = ( Kc ) 1/(x+y)
Solubility for binary electrolytes xxyy
or with x =y = 1
c·=~
Sa= ((aA):~aBY) l/(x+y)
Supersaturation S = ((CA):~CB)Yy/(x+y)
S=~= ,.;c;:c;
c· ~
~ 0.8
.£
.a
::J
0 0.4
Cl)
ll~
0
0 20 40 60 100
Temperature a [°C]
Figure 2.1. Solubility curves for several anhydrates.
Temperature a [°C]
Figure 2.2. Solubility curves for several hydrates.
line and every ternary mixture by a point inside the triangle. The area above
a saturation isotherm toward the solvent corner represents undersaturated
solutions. This type of triangular network is suitable for drawing·up mate
rial and mass balances for evaporative crystallization, as will be shown later.
As there is no clear boundary line between crystallizations from solutions
and from melts, such difficulties in distinction also arise in the discussion of
phase-state diagrams. When only one component precipitates in a pure state
from a binary, real liquid mixture, we speak of crystallization from solution.
However, when a binary mixture exhibits virtually ideal behavior, mixed
crystals precipitate during crystallization and this is known as melt crystal
lization. The terms eutectic (Latin for "well shaped") and peritectic (Latin
for "covered") will be explained in more detail by means of equilibrium
diagrams. According to Ref. 2.1, a distinction can be made between two
Physical and Chemical Properties 9
main groups of binary mixtures (see fig. 2.4): (a) eutectic systems and (b)
systems with mixed-crystal or solid-solution formation.
According to Roozeboom [2.2], noneutectic systems can be divided into
five subgroups:
Type I. A (x = 0) and B (x = 1) form a continuous series of mixed crystals
(e.g., anthracene/carbozole) [2.3].
Type II. A and B form a continuous series of mixed crystals, which is, how
ever, modified by a maximum value (e.g., o-carvoxim/L-carvoxim) [2.4].
Type III. A and B form a continuous series of mixed crystals, which is,
however, modified by a minimum value (e.g., m-chloronitrobenzene/m
fluoronitrobenzene) [2.5].
Type IV. A and B form a series of mixed crystals interrupted by peritectic
(e.g. eikosanol/hexakosanol) [2.6].
Type V. A and B form a series of mixed crystals interrupted by an eutectic
(e.g. azobenzene/azoxybenzene) [2.7].
A mixture with an eutectic concentration crystallizes when the tempera
ture falls below the eutectic temperature, forming uniform mixed crystals;
that is, the two types of mixed crystals crystallize simultaneously. During
10 Mersmann
0)
l e--------i
~
E solid
i!l!~---~--'
0 Mole fraction x
@ System with Eutecticum
&~
y .s • I
Xs &~•XX::, X
y'
X
O x,y· 0 x,y· 1
0 x,y· O x,y·
@ Systems with Mixed Crystal Formation
peritectic solidification, first one phase crystallizes, then the other. In prac
tice, eutectic systems are most common, followed by pairs of substances
exhibiting continuous formation of mixed crystals (type I). Figure 2.5
explains the systems in more detail by giving some concrete examples.
The overview panel in the figure contains some liquid-solid equilibria. The
left-hand side of the figure shows the melting-point diagram with respect to
the molar or mass fraction, and the right-hand side shows the equilibria: the
relationship between the concentration y* in the liquid phase and the con
centration x in the solid phase. The top line represents a virtually ideal
mixture exhibiting partial solubility, and the bottom line a mixture exhibiting
no solubility of its components. The solubility diagrams on the left contain a
melting-point line corresponding to concentration y* and a solidification line
corresponding to concentration x. As a system in equilibrium must have the
same temperature in both phases, the corresponding x value can be read off
for every y* value and be represented in the equilibrium diagram on the right.
If a homogeneous solution is cooled, the left diagram can be used, for exam
ple, to determine the temperature at which the first crystal is formed and the
concentration of the crystal. It is then possible to determine the concentra-
Physical and Chemical Properties 11
80 Homogenous solution
·c
40 Homogenous Solid
0 Mole fraction x,y*
Solidx
2,4 Dinitro- 2,4 Dinitro-
brombenzene chlorbenzene
100
·c
IC>
I!!
80
.3 ·=--
~
(1)
C.
60 §
E ~
i5! 40 bMixed•
0 crystals
crys
Homogenous I
1
solution
150
,c,
I!!
~ 100
(1)
·c
/ XL
§
crystals
~ ~
C.
E
i5! E
oo~----------'
Solidw
Ice KCI
tion of the liquid phase that is in equilibrium with the solid phase, the per
centages of liquid and solid, and the composition of the last drop of liquid
before it solidifies. These diagrams not only resemble boiling-point and dew
point curves of vapor and liquid equilibria but also allow analogous applica
tions (law of mixtures, lever principle).
12 Mersmann
l::!..H;* = -m [alnai]
a(l/T) P fior a;-+ a;
* (3.1)
represents the differential heat of solution, which describes the heating effect
when this solid i dissolves in a solution that is almost saturated. In such
solutions, the solid and fluid phases are practically in equilibrium, which
explains why the values t::..Hi and ai have an asterisk. As the value t::..H;
represents the heating effect of the last soluble molecules, this is known as
the last differential heat of solution. The negative value of t::..Hi represents
the heat of crystallization, which is either given off or taken up during
crystallization. First differential heat of solution is the term that applies
when a substance i is dissolved in a pure solvent or in an infinitely diluted
solution:
t::..H;* = -m [alna;]
a(l/T) P 1or
t'
a;-+ o (3.2)
The entire heat ofsolution is the energy released or absorbed when a solute is
added isobarically/isothermally to a pure solvent until a specific concentra
tion is obtained. As illustrated in Table 3.1, the first heat of solution of
different substances can vary considerably.
If the mixtures are ideal, a; = Y; or JI;= 1. Equation (3.1) then becomes
In this case, the heat of solution is equal to the heat of fusion of a pure
substance (cf. Chapter 13). Equation (3.3) can be written in the form
(3.4)
or
a(ln Ci) ~ (- t:,.Hi) (3.5)
a(ln T) ffiT
It will be shown later that the slope a(ln Ci)/a(ln T) of the solubility curve or
the heat of crystallization is very important for the decision on the mode of
crystallization and also for the problem of encrustation. Principally speak
ing, the slope of the solubility curve can be positive, zero, or negative
(inverse solubility dependence). Further information is given in Chapter 9.
Heat balances can easily be established in enthalpy-<:oncentration dia
grams for binary systems. In such diagrams, the molar enthalpy H (or the
specific enthalpy h) is plotted against the mole fraction y (or the mass frac
tion w) with isotherms as a parameter, see Figure 3.1. The molar enthalpies
of the pure substances are the product of the molar heat capacities and the
temperature: H = CPT. Most diagrams are based on the Celsius tempera
ture {}, but any convenient temperature can be chosen. If the binary systems
behave ideally, the isotherms in such enthalpy-<:oncentration diagrams are
straight lines. This means that mixing the two components that have the
same temperature does not lead to a change in temperature and the heat of
mixing is zero. However, nearly all aqueous systems show real behavior,
with the result that the heat of mixing is negative (exothermic) or positive
(endothermic). Because enthalpy-<:oncentration diagrams for real systems
can only be established by measuring the heat of mixing, only a few dia
grams have hen published in the literature.
As an example, Figure 3.1 shows the temperature-<:oncentration diagram
for a magnesium sulfate-water system. Line EB is the melting-point line (or
Physical and Chemical Properties 15
80 l
·c
60
/J
IA L
I
0 j__ _E Va H~ ____!_
I
C
0 0.1 02 03 0.4 05
kg MgSO4 / kg mixture
Mass fraction w*
100
kJ/kg
0B
·100
·200
.c.
>-
.9- ·300 L
<ti
C
~(I)
J
,g ·(00 K
·o(I)
a.
(/J -500
liquidus), and the straight line EC is the solidification line (or solidus) in the
range up to the concentration of the eutectic point E. The melting-point line
of different hydrates starts at point E and continues through DAF. In the
EBC field, solid water (ice) and magnesium sulfate solutions are in equili
brium. The isothermal ECI triangle of -3.89°C represents a three-phase
system with a magnesium sulfate solution of composition E consisting of
16 Mersmann
solid water (ice) and magnesium sulfate crystals that contain 12 molecules of
water for every molecule of MgSO 4. The DHJ triangle also represents a
three-phase system consisting of solid MgSO4·12H 2O and MgSO4•7H 2O
and a saturated magnesium sulfate solution with a mass fraction of
w* = 0.21. The EDHI and DJLA areas contain two-phase systems in
which MgSO4-12 H 2O or MgSO4-7H 2O crystals exist in saturated solutions.
More information is provided in [3.1].
Examples
-=5.=
(bond energy n
Crystal type Component Lattice forces Properties in kJ/mol) [
n·
Metallic lattice Atomic residue with Metallic bond Not very volatile, high electro Fe (400) e.
free outer electrons conductivity and thermal Na (110)
conductivity Bronze
Ionic lattice Ions Ionic bond Not very volatile, nonconductor, NaCl (750) ;·
i
r,J
'""'
...:a
18 Mersmann
1) triclinic 2) monoclinic
[llJ
4) tetragonal 5) hexagon.
(bl
6) rhombohedral 7) cubic
(cl
Figure 4.1. (a) Elementary cell (upper left); (b) explanation of Miller's
index; (c) different crystal systems.
i
to;
=
=
Physical and Chemical Properties 21
I,'
1Q)
0
0.4
C
0 0.2
0
0:::, ragonite:
"C 0
LU
0 900 1800 2700 3600
Time't [s]
Figure 5.1, in which, besides the parameters concentration and time, the
specific power input in a stirred vessel decides whether vaterite or aragonite
is formed. Another example is the precipitation of aluminum hydroxide,
which occurs in five different polymorphs and two different stoichiometries
as AlOOH or as Al(OH)J. Temperature, interfacial tension, and Zeldovich
factors are the parameters that determine which polymorph predominates;
the latter two have to be seen as specific to each polymorph and with respect
to specific adsorption (see Chapter 8) [5.1].
With respect to polymorphism, it is not only the kinetics of the chemical
reaction, nucleation, crystal growth, and phase transformation that can be
complex but also the determination of thermodynamically stable states or
phase diagrams. Very long periods of contact between a solid phase and a
liquid phase are often necessary to obtain the equilibrium state. When
equilibrium is approached, changes in phase properties become smaller
and smaller and it may be difficult to decide on the final equilibrium
state. Sometimes, slow crystallization with seeding may finally lead to
a reliable point in a phase diagram, but even this procedure can fail.
The term "dynamic phase diagram" is used to characterize this behavior.
This is a major drawback because supersaturation as the driving force for
crystallization kinetics always refers to the equilibrium state.
When crystallizing organic substances, enantiomorphs and racemates can
play an important role. Two crystals of the same substance are said to be
enantiomorphous when they are the mirror images of each other. This can
be seen in Figure 5.2 in which the L- and the n-form of amino acids are
depicted.
22 Mersmann
H~,I
Mirror plane
=~"Mi,
L.........,
+H
OH
CH,
L-Alanin D-Alanin
Figure 5.2. The L-form and the D-form of the amino acid alanine.
many imperfections [6.1]. In Figures 6.1 and 6.2, polycrystals are illustrated,
and Figure 6.3 shows the formation of polycrystals: At first, a large number
of nuclei are generated.
After the growth period 't' (distance of nuclei/growth rate), the crystals
touch each other and then grow together. Because growth is hindered, the
polyhedral shape of the single crystal is abandoned and a polycrystal with
grain boundaries is obtained. It will be shown later that the growth of such
crystals and the attrition behavior depend on these grain boundaries and on
the mosaic structure of the polycrystal. X-ray investigations of 500 µm
KAI(S04h•l2H 20 polycrystals have shown that the size of a mosaic
block is approximately 200 nm. This means that the polycrystal may consist
of 10 10 mosaic blocks. The energy Yee per unit area of grain-boundary area
can be calculated from
iifill §flli!illiiii!ll!l!lff
0§o
OOOV'V" 0 0 0 00000000000000 0
0 0°ggogooggol!!~g
00000000000000000000000808ggggggoooooogeo
0000000
000000000000000000000000000000 00000000000
8 00000000000000000000000000000000000000000000
0000000000000000000 0 00000
00000000000000000
000000000000 00
0
°
0000000000000000
oooooooooooogooooo 00 0 00000000000000000oooo
0 000000000000000000000
0000000000000 00 ~ oggooooooooooooooooooo 0 o
0000000000000 000=0 ooooOOoooooooooooooo o
0000000 0 0000 0000 gt>o 00~00000ooooooooooogooo
oooooo 0000 oao 0000 oo -.. ooo 000000000000 ooo
o 0000 00 0 000 o oooo oo 00 0000 ooo oooooooooooog
0 ooo 0000 ooo 000000000000
dislocation
Figure 6.2. Grain-boundary angle between two mosaic blocks.
~ 0 [!]
®
0 0 I!]
~
~ 0 ~
~
a) b) c)
'f/=0
f/Li;; (6.2)
In the grain boundaries, the energy YccAo, with A 0 as the surface of all
mosaic blocks, is stored, which results in an enlarged energy of the poly
crystals or chemical potential JJ,c with the consequence that the supersatura
tion b..µ, = JJ,c - µ,L is reduced. Here, the potential µ,L represents the
Physical and Chemical Properties 25
•• • • •••• •••
.. . .. .. .........·:1-
•• • ••••••••
...........
••
••
•
•
••• •••• •
• ••• • • •
•• •• •• •• •• •• •• •• •• • □ ==E.0
••••••••• : I
••••••••• ;__i_
.• .• .• .• •. •. •. •. •.
•••••••••
••• • • •• ••
••• •• ••••
•••
.. .
. . . .• .•. •••
••• •• •
-
•••
. •••
b 1}
potential of the solute in the solution. In Figure 6.5, the effective relative
supersaturation O'eff is plotted against the crystal size L with the mosaic
spread 1J as the parameter. The smaller the crystal (attrition fragments)
and the larger the mosaic spread TJ, the more the supersaturation is reduced
and can become negative (dissolution) [6.2].
:C 0.05
J
§ 0.03
~
~ 0.01
~ 0
~ -0.01
::::,
en
Cl> Potash alum
>
:;:::; -0.03 8=25°C
55
u
-0.05 +--------~~---~~~-,....../
10 100
cr=0.05
Figure 6.5. Effective relative supersaturation against the crystal size for
different mosaic spreads.
26 Mersmann
screw dislocation
...
1 ...
0
which indicates the direction and the magnitude of slip. With respect to
these defects and the grain boundaries, the tensile strength a of polycrystals
is much lower than it would be for a perfect crystal. This strength for defect
free crystals is
a~ 0.1 E (6.3)
Young's modulus E and the hardness H. Materials with a high E/H ratio
( > 180) exhibit ductile behavior (chorides, iodides, bromides, and fluorides),
whereas solids with E/H < 100 are brittle. Sulfates, phosphates, and nitrates
are often brittle. The hardness of a solid is considered to be its resistance to
local plastic deformation, which may be caused by the indentation of a
Vickers pyramid. Dividing the load applied by the area of the indentation
formed gives the Vickers hardness:
F
Hv = 1.854 did2 (7.2)
Here, Fis the force applied and d1 and d2 are the two diagonals of the plastic
indentation. The factor 1.854 results from the geometry of the indenter,
which is a pyramid having an included angle of l 36°. When the change in
geometry of the indentation during unloading is neglected, the hardness is
equivalent to the contact pressure during loading. Assuming that the contact
pressure of a plastically deformed cone and the Vickers hardness are the
same, we can use the appropriate Vickers hardness of a substance in equa
tions presented later. However, the quasi-static hardness measured in this
way can be significantly lower than the hardness measured under impact
conditions. This is because plastic flow requires the movement of disloca
tions in the crystal, which depends on the strain rate.
Because the required pressure necessary for plastic deformation is high, it
is assumed that ductility is restricted to a region limited by the area of
contact and that the material is elastic elsewhere. This regtion is taken to
be a spherical segment of the radius a below the circle of contact. The
plastically deformed region is assumed to be under hydrostatic pressure,
which coresponds to the contact pressure. This pressure is equivalent to
the dynamic hardness after complete loading. The average strain energy
per unit solid volume Wv depends on the hardness Hv, the shear modulus
µ,, and the radius a based on the radius r (see Fig. 7.1 [7.4]):
hydrostatic core
w ~3H
V
2
-v -
16µ,
(;)4
r
(7.3)
When Rittinger's law applies, this energy can be transformed into surface
energy r or fracture resistance after a crack, and the characteristic size of a
fragment in a given direction L; will depend on the strain energy per unit
volume according to
w =2-
r (7.4)
v L;K
K is an efficiency constant that is dependent on the nature of the stresses
(tension or shear or both). Up to now, it has been difficult to predict the
fracture resistance (r / K) exactly. A simple equation has been proposed by
Orowan [7.5]:
(r) = (
K l.1E nCcNA
1 ) 1 3
/
= 1.1Elo· (7.5)
(7.6)
Fe is the critical force necessary to form cracks. In Table 7.1, data for Hv,
Fe, and We are given for 10 solid substances that were investigated by
indentation tests. When the energy We required for cracks is determined
experimentally, the fracture resistance (r / K) can be calculated from
(-r) =
K
0.1
wI/3 HS/3
e
µ,
V (7.7)
As has been shown, all mechanical material properties important for crystal
attrition and deformation can be obtained from indentation tests. Gahn
[7 .6) investigated 10 inorganic and organic substances. Indentation tests
were carried out on a total of 27 different crystallographic faces; see
Table 7.2. The results are summarized in Table 7.1, where data for the
hardness Hv, the shear modulus µ,, the ratio E/Hv, the critical work We,
the density Pc, the molar mass M, the number of elements n, the length /0 ,
and the fracture resistance (r / K) are given. By comparing the hardness as
~
Hv µ E lo r/K, r/K,
[xl0-6 [x10- 9 [xl0- 9 w;, Pc M [x 10 10 [Eq. (7.5)] [Eq. (7.7)]
Substance N/m2] N/m 2] N/m 2] Ve E / H v [ x IO 10 J] [kg/m3] [kg/kmol] n m] [J/m2] [J/m2]
L( +)-Tartaric acid 1030 9.58 25.4 0.32 25 36 1759 150.1 16 2.1 8.95 16.8
Potash alum 754 7.96 20.3 0.27 27 7 1760 474.4 48 2.1 7.26 7.0
Citric acid 454 4.71 12.6 0.35 28 35 1542 210.1 24 2.1 4.54 8.7
Potassium sulfate 1502 17.4 44.1 0.27 29 30 2662 174.3 7 2.5 18.7 16.4
Magnesium sulfate 649 9.06 23.6 0.31 36 48 1680 246.5 27 2.1 8.37 9.1
Thiourea 137 3.08 8.09 0.33 59 1800 1405 76.12 8 2.2 3.08 6.7
Ammonium sulfate 355 8.90 23.4 0.32 66 41 1769 132.1 15 2.0 8.06 3.2
Potassium nitrate 265 7.17 18.9 0.32 71 59 2109 101.l 5 2.5 8.09 2.8
Sodium chloride 166 14.7 36.9 0.25 222 - 2163 58.44 2 2.8 17.7
Potassium chloride 91 9.44 24.l 0.28 265 - 1984 74.56 2 3.1 12.9
[
a
=
5
Physical and Chemical Properties 31
well as the crack patterns formed by indentation tests, the material behavior
with respect to attrition can be grouped into three classes:
------- -$-
side view top view
radial cracks
lateral cracks
Figure 7.2.
-------
--- ~
Schematic of the observed radial and lateral cracks.
32 Mersmann
Figure 7.3. Indentations on the {11 l} face of potash alum (left) and the
{101} face of citric acid (right) at a load of 0.1 N.
Physical and Chemical Properties 33
lateral
crack
{110} {010}
Figure 7.4. Illustration of crack patterns and surface pileup for indenta
tions on potassium nitrate.
Hv
a,=-
3
(7.8a)
1012
IJ 1011
Cl)
::,
:5
'tJ
0
E 1010
roQl
.c
(/)
109
0.1 0.15 0.2 0.3 0.4
lnteratomic distance [nm]
Hv
a1
=~ +
3
[1 ln(0.4£)]
3a 1
(7.8b)
.!!}__(r)
H~ K
3
(7.11)
I
O; 10
11
+---1------4
ti)
:::,
"3
109 +---,i..-----,~----1----+...
107 108 109 1010
Vickers hardness Hv [Nim']
,, plastic
10
crystals
.s
ti;
1010
Q)
0
+AgC~
C: •~~~~o"nite)
sion of solids. In Figure 7.7, the abrasion resistance as the reciprocal of the
abraded volume is plotted against the Vickers hardness for many crystalline
solids. Most of the crystals are in the semibrittle region. The figure clearly
shows that the abraded volume is not a function of the hardness alone. Even
the morphology plays a role, which can be seen from the different behavior
of calcite and aragonites as modifications of calcium carbonate.
8TCL
•CL= 'YCL+A 8A (8.1)
For the first term in equation (8.1), which contributes to the surface stress,
the interfacial tension, 'YCL, the solid molecules can be assumed to be fully
mobile, with every deviation from this assumption being described in the
second term. 'YCL is defined as a material-specific property; however, it is still
fixed to a single face of the crystal and the overall or global interfacial
tension of the crystal collective will also depend on the morphology and
the structure of the crystals. A material will exhibit a different global inter
facial tension when it crystallizes in needles then it will exhibit when it
crystallizes in cubes or plates. This is because of the varying extension of
the different crystal faces, where each of the faces has a specific interfacial
tension. The question now arises as to what the physical understanding of a
global interfacial tension should be, which also depends on the geometry of
the crystals. However, it has to be kept in mind that, so far, it has not been
possible to measure the interfacial tension of a single crystal face directly
and reliably. Almost all experimental data for 'YCL result from measurements
of rates of primary nucleation, expressing, in fact, a global interfacial ten
sion of a crystal collective. On the other hand, for the application in models
to describe the kinetics of the crystal collective in an industrial crystallizer, a
C*/Cc
f 40 ..---o~.a_ _o~.5_ _~10_·1_1~0_·2 _~10_·•-10~·10....,,
~ ;i
cJ
S2"
·~
II 10
C 5
0
"iii
C
~
Iii
~
~ 0.5
K=0.333
""" ,'
,,,,',,'
/
,,,,'
fil ,,
c / / "K = 0.414
.Q
rn
C
Q) o.1..,.__ _- , . - ~ - - - ~ - , . - - - - 1
E 0.1 0.5 1 5 10 40
i5
Concentration Ratio In (CJC*)
YCLd;,
kT
= 0.272 . In (Cc)
C* (8.3)
L• YCLVm
crit = 4 kTlnS (8.4)
dm
3
= 67r dm,s
3
(8.5)
lyte solutions [8.8, 8.9]. This can be applied more generally to all crystal
systems, taking into account that with the dissociation of the lattice mole
cules, the interface is not necessarily in a stoichiometric composition.
Depending on the concentration of the dissociated ions in the mother liquor,
there is generally a surplus adsorption of one lattice ion above the other. In
addition, other substituting ions may be adsorbed directly on a surface
lattice site and thus affect the interfacial tension. This may be included in
a physical model derived from Eble and Mersmann [8.10]. The specific
adsorption of ions leads to an excess charge, which results in a Coulomb
potential that, in turn, attracts oppositely charged ions form the bulk of the
solution. The surface charge density, a, as the excess adsorption of potential
determining ions of valence, z, in the electrical double layer is
a= zF(f'A+ - f'B-) (8.7)
with Fas the Faraday constant. f' A+ and f' B- are the surface densities of the
lattice ions A+ and B-, respectively, in the adsorption layer. The maximum
loading and thus the maximum surface charge may be assumed to be equal
to the total number of lattice sites at the interface, given by the molar crystal
density Cc with the length of the elementary cell a0 orthogonal to the inter
face:
(8.8)
With the Gibbs adsorption equation, the decrease in interfacial tension with
the specific adsorption (taking into account the nonstoichiometric case
f' A+ =I= f'B-, at constant temperature and pressure and in the case of equili
brium throughout the two-phase system) is given by
(8.9)
According to the considerations of Gilbert and Rideal [8.11 ], the differential
of the electrochemical potential aµ,A+ can be split up into an electrostatic
part, which is necessary for making an ion approach the charged surface,
and a chemical part that is due to the sorption described here using the
Langmuir equation. Combined with equation (8.7), the decrease in the
interfacial tension due to the specific adsorption of an ion of valence z
then reads
(8.10)
between the surface charge a, the ionic strength/, the Debye length 1/K, and
the electostatic potential % is given by [8.12, 8.13]
a= 4 FI sinh(F%) (8.11)
K 291T
The integration of equation (8.10) thus leads to the general equation for the
interfacial tension YCL,,,, which is valid for a surface having the surface
I
charge density a:
-
N""
E 0.0 0.2 0.4 0.6 0.8 1.0
3 ., 0.00
~
<l
Cl) -0.02 ,10•,
10-•,
~0- <ll'J~'
Ill
(II
~
u -0.04
Cl)
"C ,o~o
tt>
C:
0 ~v
'iii -0.06 ,o
.s
C:
i5
·u -0.08
~ calculated for:
.s
.E -0.10
a== 1C/m2
Figure 8.2. Reduction in the interfacial tension l:!.yCL with the surface
charge density. Calculated from equation (8.12), valid for amax = l C/m2 •
(From Ref. 8.10.)
42 Mersmann
a low ionic force makes the interfacial tension sensitive to the sorption of
potential determining ions, the effect being essential at high ratios of a/amax
together with high values of amax.
REFERENCES
[2.1] S. Rittner and R. Steiner, Die Schmelzkristallisation von organischen
Stoffen und ihre grosstechnische Anwendung, Chem. Ing. Techn.,
57(2): 91 (1985).
[2.2] H. W. B. Roozeboom, Die Loslichkeitskurve fiir Salzpaare, welche
sowohl Doppelsalz als Mischkrystalle bilden, speziell fiir Salmiak mit
Eisenchlorid, Z. Phys. Chem., JO: 145 (1892).
[2.3] E. Funakubo and S. Nakada, Jpn. Coal Tar., 201 (1926).
[2.4] H. Reinhold and M. Kircheisen, Eine Methode zur Untersuchung
biniirer Systeme. 4. Mitteilung: Das "Auftauschmelzdiagramm" als
Mikromethode, J. Prakr. Chem., 113: 203 (1926).
[2.5] M. Hasselblatt, Uber die lineare Kristallisationsgeschwindigkeit iso
morpher Mischungen, Z. Phys. Chem., 83: 1 (1913).
[2.6] H. Schildknecht, Zonenschmelzen, Verlag Chemie, Weinheim, p. 27
(1964).
[2.7] W. Polaczkowa, Preparatyka organiczna, Warsaw, p. 51 (1954).
[3.1] International Critical Tables (E. W. Washburn, ed.), McGraw-Hill,
New York, Vol. IV, (1933)
[4.1] 0. Sohnel and J. Garside, Precipitation, Butterworth-Heinemann,
Oxford (1992).
[5. l] A. Eble and A. Mersmann, The kinetics of boehmite (y-AIOOH)
crystal growth, nucleation and aluminium hydroxide polymorphism,
J. Cryst. Growth (in press).
[5.2] J. Franke, Uber den Einfluss der Prozessparameter auf die
Fiillungskristallisation am Beispiel von Calciumcarbonat und
Calcium sulfat Dihydrat, Thesis, TU Miinchen, 1994.
[6.1] J. Mullin, Crystallization, Butterworth-Heinemann, Oxford (1993).
[6.2] U. Zacher, Die Kristallwachstumsdispersion in einem kontinuierli
chen Suspensionskristallisator, Thesis, TU Miinchen (1995).
[7.1] A. Reuss, Berechnung der Fliepgrenze von Mischkristallen auf
Grund der Plastizitiitsbedingung fiir Einkristalle, Z. Angew. Math.
Mech., 9: 49 (1929).
[7.2] R. Hill, The elastic behaviour of a crystalline aggregate, Proc. Phys.
Soc. London., A65: 349 (1952).
[7.3] Landolt Btknstein, New Series, Springer-Verlag, Berlin (1966/1977).
Physical and Chemical Properties 43
Crystals are created when nuclei are formed and then grow. The kinetic
processes of nucleation and crystal growth require supersaturation, which
can generally be obtained by a change in temperature (cooling in the case of
a positive gradient dC* / dfJ of the solubility curve or heating in the case of a
negative gradient), by removing the solvent (usually by evaporation), or by
adding a drowning-out agent or reaction partners. The system then attempts
to achieve thermodynamic equilibrium through nucleation and the growth
of nuclei. If a solution contains neither solid foreign particles nor crystals of
its own type, nuclei can be formed only by homogeneous nucleation. If for
eign particles are present, nucleation is facilitated and the process is known
as heterogeneous nucleation. Both homogeneous and heterogeneous nuclea
tion take place in the absence of solution-own crystals and are collectively
known as primary nucleation. This occurs when a specific supersaturation,
known as the metastable supersaturation Ll.Cmet, is obtained in the system.
However, in semicommercial and industrial crystallizers, it has often been
observed that nuclei occur even at a very low supersaturation Ll.C < 6.Cmet
when solution-own crystals are present (e.g., in the form of attrition
45
46 Mersmann, Heyer, and Eble
Homogeneous
Primary
Heterogeneous
Nucleation Contact
Shear
Secondary Fracture
Attrition
Needle
.u /
// I
I I
1
/ // /
c5 ,/
;/'
/
,,~ /
/
/
/
/
C:
0 ,./ ,/' , / Metastable zone
,,-· ,,, ., width for nucleation
~C: - secondary
~ -primary,
C: heterogeneous
0
u -primary,
homogeneous
Temperature T
or, equivalently, the extent of penetration into the metastable zone is dif
ferent for each process due to different physical mechanisms. Homogeneous
nucleation is treated in Sec. l, including discussions on classical theory (Sec.
I.I) and kinetic theory (Sec. 1.2), and rules for industrial application are
given in Sec. 1.3. In Secs. 2 and 3, heterogeneous and secondary nucleation
are presented, and, finally, all three mechanisms are compared in Sec. 4.
1. HOMOGENEOUS NUCLEATION
1.1. Classical Theory
The classical nucleation theory dates back to the work of Volmer and Weber
[l.l, l.2], who were the first to argue that the nucleation rate should depend
exponentially on the reversible work of the formation of a critical cluster
and was later extended by authors such as Becker and Doring [l.3], Farkas
[1.4], Zeldovich [1.5], Frenkel [1.6], and others [1.7]. In order for a new phase
to appear, an interface must be formed, which (in the absence of impurities
or suspended foreign material) occurs by small embryos in the new phase
being formed within the bulk metastable phase. These embryos are formed
due to spontaneous density or composition fluctuations, which may also
result in the spontaneous destruction of such an embryo. The creation of
nuclei can, therefore, be described by a successive addition of units A
according to the formation scheme
k,4
Ai+ A = A2; A2 +A= A3; ... ; An+ A<=> An+I
kD
(1.1)
(l.5)
Activated Nucleation 49
'\.'\.
103 '\. '\. dm =0.7 nm
-..] 10
E
'O
2
0
~
a:
10
_,O O 8-=20DC
10·3 10·2 10-1 100 101
Supersaturation In S
Figure 1.2. Ratio Le/dm with respect to the natural logarithm of super
saturation S for different interfacial tensions and molecule diameters.
In the case of such nuclei, the rate constant kA of addition is as great as that
of decay kn. Therefore, neither growth nor decomposition takes place. The
last four equations yield the following relationship for the critical nucleus
diameter L~rit in the case of spherical nuclei:
Because the free enthalpy !:::,.G for nucleus sizes L > Le decreases with the
nucleus size, the addition reaction takes place by itself due to the laws for
disturbed equilibria (i.e., the nucleus can continue growing). On the other
hand, in the range L < Le, the change in free enthalpy increases with
50 Mersmann, Heyer, and Eble
increasing nucleus size. This means that the rate constant of decay is greater
than that of growth, and the nucleus disintegrates.
The rate of primary homogeneous nucleation according to the classical
nucleation theory can be obtained by calculating the number of critical
clusters that cross the nucleation barrier described earlier and is then
given by
(1.9)
It can be seen that the rate is a product of three factors:
• ne, the number concentration of critical clusters
• k, the rate at which clusters cross the barrier
• Z, the imbalance factor
all of which will be explained in more detail.
Assuming that the cluster size distribution n; is caused by random
collisions of molecules and can be described by a Boltzmann distribution,
we obtain
n; ky'
= n0 exp ( - l),.G.) (1.10)
(1.11)
which can be achieved by inserting Le [Eq. (1.8)] into the expression for l),.G
[Eq. (1.2)]. Equation (1.6) for the critical cluster diameter Le = J Ae/rr gives
the following expression for the distribution of critical clusters with a = 1:
(1.13)
For nucleation from the gas phase, the rate at which monomers impinge
successfully on the surface of a nucleus of size L is given by
w
k=Anq- (1.14)
4
Here, A is the surface area of a nucleus, n is the number concentration of
monomers, w is the mean velocity of monomers, and q the "condensation"
coefficient, which is the fraction of monomers hitting the surface that actu
ally stick. Assuming that every monomer that hits the surface is incorpo
rated (i.e., q = l) and that the number concentration of monomers is hardly
not reduced by the nucleation process, for an ideal gas
z = ✓2rr~T (l.17)
where
(l.18)
is the second derivative of the free energy with respect to cluster number i at
the top of the barrier. Using the expression of the classical nucleation theory
[Eq. (l.2)] for !).G and assuming that the nucleation process is volume con
serving, leading to
(l.19)
as the number of elementary units of a cluster with diameter Le, for the
imbalance factor Z
52 Mersmann, Heyer, and Eble
(1.20)
is obtained. Inserting the values of nc [Eq. (1.13)], Z [Eq. (1.20)], and k [Eq.
(1.16)] into the expression for Bhom [Eq. (1.9)] finally gives the following
form of Bhom:
~ 1
Bhom -
_ (
l.5DAB CNA
)7/3
VkT CcNA
(1.21)
16 YCL 3 1 1 ]
( ) 2
x exp [ -3n(kT) CcNA (vlnS) 2
The derivation of this surface tension and especially of the factor K are
elucidated in Chapter I.
It should be noted that for the derivation of the enthalpy of a critical
nucleus according to equations (1.2) and (1.12) and thus for the rate of
homogeneous nucleation, it was assumed that the nucleus consists of only
one component and that no solvent molecules are incorporated into the
lattice; that is, the solvent is insoluble in the solid phase. This, however, is
not necessarily the case in every nucleation process and there might be some
two-component or multicomponent nucleating systems. Although it is
beyond the scope of this book to go into more detail with regard to these
special processes, we wish to draw the reader's attention to the fact that
there are some differences for multicomponent systems. The main difference
is the evaluation of the free enthalpy of formation because one has to take
into account the ratio in which the different molecules are incorporated into
the cluster and how they may affect each other. In addition, the preexpo
nential term of the rate of homogeneous nucleation may be altered due to
changes in the diffusivities of the different types of molecules, which not
only act on each other within the solid phase but also in the solution.
Besides the nucleation of multicomponent systems, the classical nuclea
tion theory presented is strictly speaking not valid for the case of nucleation
in the presence of temperature changes, because its derivation is based on an
Activated Nucleation 53
barrier due to the cluster and its surrounding fluid. Using the mean disso
ciation time r(n), the rate at which molecules abandon the cluster can be
written
Ns(n)
ot.=-=- (1.23)
r(n)
where Ns is the number of molecules in the surface layer (i.e., within a
distance R < r < R + dm of the center of the cluster). Here, R is the radius
of the cluster and dm is the diameter of a molecule. If r > R + dm, the
molecule dissociates from the cluster.
If the cluster is a uniform spherical solid, the number of molecules within
it is given by
(1.24)
which is equivalent to equation (1.19). The number of surface molecules is
then, according to Ruckenstein and Nowakowski [1.30],
(1.25)
J~
g {J(g) - a(g)
= Jo {J(g) + a(g) dg
{J(l)n0
w(g) (1.26)
Bhom,kin =2 exp[-2w(g)] dg'
Bhom,kin = (0.5)tJ(l)no (w
"(: *))0.5 exp[2w(g*)] (1.27)
(1.28)
56 Mersmann, Heyer, and Eble
or, in general,
(1.29)
The value C* /Cc now represents dimensionless solubility and the expression
Bhom = Bhom/ DAB(NACc) 513 is the dimensionless rate of primary, homo
geneous nucleation. If a certain dimensionless rate Bhom is given, the dimen
sionless metastable supersaturation (ll.Cmet)/Cc of homogeneous primary
nucleation can be calculated for a system with C* /Cc:
.
-
~
::::, .._1:1
10--
<ti
~
Q)
/ / ~
a.
::::, / / .._1:1
en v= ;/ / / / / •
o// .._1:1
(/J
(/J 10"'
Q)
C: / / / / ~
.._1:1
0
"ii)
C:
/ / / / //
E
Q)
10.. /
i5 10.. 10.. 10"' 10-- 10°
Dimensionless Solubility C*/C 0
values for Cc and DAB• Note the strong influence of the stoichiometric
coefficient v for sparingly soluble systems [1.32].
It is somewhat surprising that all curves pass through a maximum for a
dimensionless solubility of approximately C* /Cc:::::: 0.16. The reason for this
may be that equation (1.22) is a simplified version of the general equation
(see Ref. 1.10):
(1.31)
However, these relationships are valid only when the solution is abso
lutely free of foreign particles, which is never the case in industrial crystal
lizers. Moreover, heterogeneous nucleation of all apparatus and machine
parts wetted by the solution must also be avoided, which is very difficult
and, again, not the case in practice. Besides, one also has to keep in mind
that other effects such as mixing might have a huge influence on nucleation
rates in industrial crystallizers. So far, mixing has not been considered at all,
58 Mersmann, Heyer, and Eble
2. HETEROGENEOUS NUCLEATION
2.1. Theoretical Approach
Up to now, it has been assumed that collisions between units in a solution
lead to clusters of varying size and that nuclei result from the fact that
clusters above the critical nucleus size L~rit can continue growing into crys
tals, assuming that the original solution is entirely pure (i.e., free of solid
particles). In reality, it is impossible to remove completely all solid matter
from a solvent or solution. In bidistilled water, many Si02 particles in the
size range of only a few nanometers can be found. Schubert [2. I] came to the
conclusion that the volumetric surface of these foreign particles is approxi
mately aror ~ 2.5 x 103 m2 /m 3; but how do such small foreign particles
(such as sand, rust, etc.) affect nucleation? This is illustrated in Figure
2.1, which shows a foreign particle in a supersaturated solution.
According to Blander and Katz [2.2] and also Schubert [2.1], the rate of
heterogeneous nucleation is proportional to the volumetric surface aror of
Activated Nucleation 59
foreign particles and not to the number of those particles. The model pro
posed by Schubert [2.1] describes the rate of heterogeneous nucleation Bhet
as a product of the volumetric surface aror of foreign particles and the rate
Bhet,surf of heterogeneous nucleation based on the surface of foreign particles
according to equation (2.1):
(2.1)
The rate Bhet,surf can be described as being equivalent to the rate of homo
geneous nucleation by
(2.2)
Again, the factor Z takes into account the difference between the equili
brium and stationary state, having the same definition as in homogeneous
nucleation [see Eqs. (1.17) and (1.18)]. The number of critical nuclei, nc, is
now based on the surface of foreign particles and not on the volume, as in
homogeneous nucleation, and the rate at which clusters cross the barrier, k,
takes into account not only impingements from the volume but also those
from molecules adsorbed on the surface of the foreign particle. Because the
two factors in equation (2.2), k and nc, are different than those in classical
homogeneous nucleation theory, they will be discussed in more detail.
The number nc of critical clusters depends on the number of molecules
CadNA adsorbed on the foreign surface and the nucleation energy .6.Gc,het of
heterogeneous nucleation according to a Boltzmann distribution:
(2.3)
Assuming now that the thickness of the adsorption layer does not exceed the
diameter dm of a molecule, which is valid for precipitation because of very
low concentrations in the solution, the surface-based concentration Cad can
be expressed in terms of the volume-related concentration Cad,v by
(2.4)
60 Mersmann, Heyer, and Eble
Taking into consideration the fact that the adsorption isotherm, which
describes the relation between the concentration Cad,v of units adsorbed
on the foreign surface and the concentration C of these units in the bulk
of the solution, can be described by a linear relationship due to the low
concentrations, expression
(2.5)
f = (2 + cos0)(1 - cos0)2
(2.8)
4
If the foreign surface is not flat, as in the case of nano-sized foreign particles,
the size of the foreign material has to be taken into account too, because it
influences the ratio of the truncated nucleus volume to that of a sphere of
equal size.
In Figure 2.2, the factor f is plotted against the angle 0 for the case of
a smooth, flat foreign surface. This angle ranges from 0° to 180°, depend
ing on the "wetting" of the foreign particle by the units. A contact angle
of 180° (point contact) corresponds to nonwettability or homogeneous
nucleation with f = 1. When the angle 0 is between 0° and 180°, the
nucleation work is reduced by the "wetting" surface of the foreign
particle. If 0 ➔ 0°, the foreign particle is completely wetted and both the
nucleation work and the formation of heterogeneous nuclei tend toward
zero. This shows that the better the wetting, the smaller the volume of the
truncated nucleus and the smaller the free-energy cost associated with its
formation.
Activated Nucleation 61
0 0
0 90° 180°
Contact angle 8
Figure 2.2. Factor f as a function of the contact angle.
A.=21t r.2(1-cos 0)
U0 =21t r0 sin 0
Figure 2.3. Schematic diagram of a nucleus on a solid surface.
(2.9)
The impact rate of units arriving from the volume can be described by a
model already presented for units hitting a homogeneous nucleus:
(2.10)
,isurf can be derived from the laws of statistical thermodynamics for a two
dimensional problem [2.l]:
(2.11)
1 ✓11:kTNA (2.12)
Dsurr = 4dm Cad N A 2M-
. ( 2MyCL ) .
Uc= 211:rcSlil 0 = 217: PcVRTlnSa sm 0 (2.14)
x exp[-1G)1r~~ ~]
is also the case if the adsorption constant Head or the surface diffusion
coefficient Dsurf is small.
The rate of heterogeneous nucleation depends on the same parameters as
the rate of homogeneous nucleation (i.e., the supersaturation Sa, the con
centration C, the surface tension YCL, the diffusivity DAB, the equilibrium
concentration C*, and the molar density Cc) and also on the following,
additional parameters:
• The volumetric surface aror of foreign particles present in the solution
• The adsorption constant Head
• The contact angle 0 or the factor f, which is a function of the contact
angle
• The surface diffusion coefficient Dsurr
We are not yet able to predict the contact angle 0 and the adsorption con
stant Head, but some experiments have shown that the contact angle in
aqueous systems is often in the range 40° < 0 < 53° or 0.038 <f < 0.1,
respectively. Schubert [2.1] carried out precipitation experiments with the
reactants Ba(OH)i and H 2SO4 in the presence of SiO2, Al 2O3, and TiO 2 in
the nanometer range. The particles were added in order to investigate their
influence on the rate of heterogeneous nucleation. The volumetric surface
aror of the foreign particles was varied between 5 x 103 and 2.5 x 105 m2 /m 3 •
In all cases, the rate of heterogeneous nucleation was proportional to the
surface aror of foreign particles. When aror is smaller than 104 m2 /m3 , it is
necessary to take into account the basic level of Si02 f.articles that is always
present, even in ultrapure water (aror ~ 2500 m2 /m ). It is important to
mention that it is not the number of foreign particles but their volumetric
surface aror that is the decisive parameter for heterogeneous nucleation.
2.2. Application
In Sec. 2.1, the theoretical background of heterogeneous nucleation was
discussed. Comparison with experimental results is a good tool not only
for evaluating this theory but also for providing a set of rules on how to
control nucleation and, therefore, the final product quality.
When carrying out nucleation experiments, it is expedient to plot the
nucleation rate against 1/(ln Saf In Figure 2.4, the rate of heterogeneous
nucleation, Bhet> is plotted versus 1/(ln Sa)2 for BaSO4 and SiO2 particles
(Fig. 2.4a), TiO2 particles (Fig. 2.4b), and AI2 O 3 particles (Fig. 2.4c). In
all cases, the volumetric surface of the added foreign particles was
aror = 1.5 x 105 m2 /m 3 • Additional experiments were carried out without
the addition of foreign particles. The figures show that the nucleation rate
increases with the surface aror for a given supersaturation. When the critical
64 Mersmann, Heyer, and Eble
~ [1/(m's)]
&C:
10"
0
~
Q)
0
::::, 10"
z a)
0.02 0.04 0.06 0.08
1/(ln S.)'
Supersaturation S,
10" 1000
~~~~-~-~-~
200 100 50
cc [1/(m's))
*
0:::
C:
.Q
10"
iii
Q)
0::::,
10"
z
1012 1--~~--~~----- b)
0.02 0.04 0.06 0.08
1/(ln S.)'
Supersaturation S"-
1000 200 100 ::,0
10" , - - ~ ~ ~ - ~ - ~ - - ,
cc [1/(m's)]
iC: 1015
0
~
Q)
0::::,
10"
z
1012 +------~--------- c)
0.02 0.04 0.06 0.08
1/(ln s.)'
1.2-10·8 -------------~
-0
co'°
:::c 1·10"8
0
C:
0
e-0
~
8•10·9 0
1/)
:lt
-
0
C:
~C:
0 0
(.)
0
S
0 = 138
10 40 70 100
Temperature 3 [ C] 0
- /
~
:::,
ro
10-2 / / .._<::I
~
Q) .._<S
0..
:::,
en / /
Cl)
Cl)
10.. / / / ►
.._<::I
/ / / /
-
Q)
-6
"E / / / / /
~
0 ad ~
·0
C: / / / / / / / /
Q)
E 10•
i5 10.. 10.. 10.. 10.. 10°
Dimensionless Solubility C*/Cc
In the range of high solubility (C* / Cc > 0.01 ), the metastable zone width
for the rate Bhet = 10 12 m- 3s- 1 is scarcely affected by the parameter aror and
Head· The smaller the factor f or the contact angle 0, the more the
metastable zone width is reduced. With sparingly soluble substances
Activated Nucleation 67
(.)
(J
....~ ,s .._,s ....~ cs ~ ....~ ~"
u<I 10° "' "' "' 6
// // / ~
/
C .._c::i
0
-
~::::, '
.._c::i
( ti
104
~ ~
Q) ....'s)
C.
::::,
en ►
fl)
10.. .._c::i
fl)
Q)
"E
0
/ ~
.._c::i
"in
C: /
Q)
E 10.. /
i5 10.. 10.. 10.. 10.. 10°
(C* /Cc< 10-5), the parameters aror, Head, and 0 have a strong influence on
!),.Cmet,het!Cc. For the nucleation rate Bhet = 10 12 m- 3s- 1, the metastable
zone width decrease with increasing values of aror and Head but decreasing
factors f or contact angle 0. Because these parameters are not known for a
real solution, the metastable zone width is unpredictable for substances with
a very low solubility. On the other hand, the equations may be useful for the
seeding of solutions in order to obtain controlled nucleation. This can be
done by the addition of foreign particles which can be characterized by the
parameters aror, Head, and/. When these added foreign particles are domi
nant with respect to nucleation in comparison to particles that are already in
the solution, the nucleation process can be controlled. In combination with
controlled growth, it would be possible to produce crystals with a desired
specification.
y;a2
V In Smet,s = T( kT~G
max,s
(3.1)
~Gmax s =
---·- Y~Ld!
T(-----',,2...:.=.-"-'--- (3.2)
kT (kT) v In Smet,s
Again, the expression
is introduced for the interfacial tension YCL· Combining the last two equa
tions gives
~Gmax s [Kln(Cc/C*)] 2
--~· ~ 1 f - - - - - - (3.4)
kT vlnSmet,s
The rate of surface nucleation, Bs, in nuclei per square meter crystal surface
and per second depends on the diffusivity, DAB, and the nucleation energy,
~Gmax,s, according to [3.9, 3.10] (compare with page 73)
B =~
D exp ( - ~Gmax,s) (3.5)
s d! kT
or
DAB
B ~--exp [Kln(CcfC*)J2)
( -T (------ (3.6)
s d! V In Smet,s