You are on page 1of 170

ALUMINIUM SCRAP REFINING WITH FRACTIONAL LAYER

CRYSTALLIZATION

Proefschrift

ter verkrijging van de graad van doctor


aan de Technische Universiteit Delft,
op gezag van de Rector Magnificus prof. dr. ir. J.T. Fokkema
voorzitter van het College voor Promoties,
in het openbaar te verdedigen op Dinsdag 21 November 2006 om 15:00 uur

door

Bedri DRINI

Doctorandus in de scheikunde

geboren te Dubov e Madhe te Kosovo

.
Dit proefschrift is goedgekeurd door de promotoren:
Prof. ir. L. Katgerman
Prof. dr. ir. P.J. Jansens

Samenstelling promotiecommissie:
Rector Magnificusvoorzitter
Prof. ir. L. Katgerman
Prof. dr. ir. P.J. Jansens
Prof. dr.-ing. O.S.L. Bruinsma
Prof. dr. ir. A.H.M. Verkooijen
Prof. dr. ir. L. Kestens
Dr. ir, D. Verdoes
Dr. ir. W. Boender
Prof. dr. ir. R. Boom

ISBN 9951-500-00-5
Publisher: DCE- Drini Consulting & Engineering
Web Site: www.dce-group.com

Cover page: Photo of river Drini spring, Kosova

Voorziter
Technische Universiteit Delft, promotor
Technische Universiteit Delft, promotor
Potchefstroom University
Technische Universiteit Delft
Technische Universiteit Delft
TNO-MEP, Apeldoorn
Corus, IJmuiden
Technische Universiteit Delft (rezerv lid)

PRINDERVE T MI
[TO MY PARENTS]

Stellingen

Belonging to the thesis


ALUMINIUM SCRAP REFINING WITH FRACTIONAL LAYER CRYSTALLISATION
Bedri DRINI

1. With an increase of the aluminium use in car manufacturing a sustainable recycling


technological process based on fractional crystallization is imperative.
2. The technique of fractional crystallization is much more advanced for organic compounds than
for metals.
3. Zone melting is a relatively simple and useful experimental technique to assess the ability to
refine with fractional crystallization complicated alloying systems.
4. Mother Nature taught us the principle of fractional crystallization by creating icebergs with
fresh water from salted ocean water.
5. Refining aluminium by means of layer growth fractional crystallisation is technically a viable
option in developing a sustainable technology for recycling of a high-grade aluminium from
scrap.
6. Post-purification processes, such as sweating, save time, costs and energy.
7. Theoretical analysis of the falling film crystallization process for aluminium refining show that
the same purification effect can be achieved with growth rates (i.e. production rate) of 10 times
higher than for static process.
8. People are hungry for love!
Gongje Bojaxhiu alias Mother Theresa
9. A man without a home-country is like a child without parents.
10. Giving people of the developing countries access to top education in developed world is the
best contribution to world development.

Stellingen
Behorende bij het proefschrift
ALUMINIUM SCHROOT OPWERKING MET FRACTIONERENDE LAAG
KRISTALLISATIE
1. Met een toename van het gebruik van aluminium in de autoproductie is een
duurzaam technologisch terugwinningsproces, gebaseerd op fractionerende
kristallisatie, noodzakelijk.
2. De techniek van fractionerende kristallisatie is veel geavanceerder voor
organische stoffen dan voor metalen.
3. Zonesmelten is een relatief eenvoudige en bruikbare experimentele methode
voor het vaststellen van de mogelijkheid om gecompliceerde
legeringssystemen op te werken met fractionerende kristallisatie.
4. Moeder Natuur heeft ons het principe van fractionerende kristallisatie geleerd
door ijsbergen te creren met zoet water uit zout zeewater.
5. Opwerking van aluminium door middel van laag-groei fractionerende
kristallisatie is technisch een haalbare optie voor het ontwikkelen van een
duurzame technologie voor de terugwinning van hoogwaardig aluminium uit
schroot.
6. Post-zuiveringsprocessen, zoals zweten, besparen tijd, kosten en energie.
7. Theoretische analyse van de vallende-film kristallisatie voor aluminium
opwerking toont aan dat hetzelfde zuiveringseffect kan worden gehaald met
groeisnelheden (dus productiesnelheden) die 10 maal groter zijn dan voor het
statische proces.
8. Mensen zijn hongerig naar liefde!
Gongje Bojaxhiu alias Moeder Theresa
9. Een man zonder thuisland is als een kind zonder ouders.
10. Mensen uit ontwikkelingslanden toegang geven tot toponderwijs in de
ontwikkelde wereld is de beste bijdrage aan de ontwikkeling van de wereld.

CHAPTER I .................................................................................................................... 1
1. METAL REFINING WITH FRACTIONAL CRYSTALLIZATION: STATEOF-THE-ART AND FUTURE PROSPECTS ............................................................. 1
Abstract......................................................................................................................... 1
1.0. Introduction ........................................................................................................... 2
1.1. Theory.................................................................................................................... 3
1.2. Layer based processes ........................................................................................... 5
1.2.1. Heinzer et al. Crystallizer............................................................................... 6
1.2.2. Pechiney crystallizer....................................................................................... 7
1.3. Suspension based Processes .................................................................................. 8
1.3.1. Alcoa crystallizer............................................................................................ 8
1.3.2. Yunnan Crystallizer ...................................................................................... 10
1.4. Emerging technologies ........................................................................................ 11
1.4. A look to the future.............................................................................................. 15
1.5. Conclusions ......................................................................................................... 19
1.6. Acknowledgements ............................................................................................. 20
1.7. References ........................................................................................................... 20
CHAPTER II ................................................................................................................ 25
2. SCOPE........................................................................................................................ 25
2.1. Fractional crystallization principle and aluminium refining .............................. 25
2.2. Selection of alloy systems to be studied............................................................... 28
2.3. Research goals..................................................................................................... 30
2.4. Thesis outline....................................................................................................... 30
2.5.

References....................................................................................................... 32

CHAPTER III ............................................................................................................... 33


3. SOLUTE REDISTRIBUTION DURING DIRECTIONAL
CRYSTALLIZATION AND INTERFACE MORPHOLOGY................................ 33
3.1. Introduction ......................................................................................................... 33
3.2. Solute redistribution during crystallization under equilibrium conditions ......... 34
3.3. Solute redistribution while no diffusion in the solid............................................ 36
3.3.1

Complete mixing in the liquid................................................................. 36

3.3.2. Stagnant liquid.............................................................................................. 37


3.3.3. Partial mixing in the liquid........................................................................... 40

3.3.4. Concluding remarks ..................................................................................... 43


3.4. Crystal - Melt interface morphology ................................................................... 44
3.4.1. Constitutional supercooling theory .............................................................. 44
3.4.2. Interface stability theory............................................................................... 47
3.5. Cell structure and solute redistribution in cellular solidification....................... 50
3.6. Selected models used to describe experimentally achieved solute redistribution in
directional crystallization........................................................................................... 54
3.6.1. Cheng, Irving and Kyle model...................................................................... 54
3.6.2. Impurity inclusions in crystal layer .............................................................. 55
3.6.3. Post purification steps .................................................................................. 57
3.7. General conclusions ............................................................................................ 60
3.8. References............................................................................................................ 61
CHAPTER IV ............................................................................................................... 63
4. ZONE MELTING INVESTIGATION OF THE ALUMINIUM SCRAP
REFINING POTENTIAL IN FRACTIONAL CRYSTALLIZATION .................. 63
Abstract....................................................................................................................... 63
4.1. Introduction ......................................................................................................... 64
4.2. Theoretical Framework ....................................................................................... 65
4.3. Experimental Procedure ...................................................................................... 69
4.4. Results ................................................................................................................. 71
4.4.1. Chemical Analyses........................................................................................ 71
4.2. Microscopic examinations............................................................................... 73
4.4.3. Electron Probe Micro-Analysis .................................................................... 77
4.5. Discussion............................................................................................................ 78
4.5.1. Comparison of the chemical analyses with equilibrium predictions............ 78
4.5.2. Comparison of results for AlSi6 alloy with non equilibrium BPS-model..... 81
4.6. Conclusions ......................................................................................................... 83
4.8. References ........................................................................................................... 85
CHAPTER V .................................................................................................................. 87
5. ALUMINIUM SCRAP REFINING WITH SOLID LAYER FRACTIONAL
CRYSTALLIZATION ................................................................................................... 87
Abstract........................................................................................................................... 87
5.1. Introduction ............................................................................................................. 88
5.1. Introduction ............................................................................................................. 88

5.2. Theory...................................................................................................................... 89
5.3. Experimental............................................................................................................ 91
5.4. Results and discussion ............................................................................................. 94
5.4.1 Chemical analyses ......................................................................................... 94
5.4.2. Microscopic Examinations ........................................................................... 97
5.4.3. Line scans ................................................................................................... 100
5.4.4. Inter-metallic compounds ........................................................................... 104
5.5. Conclusions ........................................................................................................... 107
5.6. Acknowledgements ............................................................................................... 107
5.7. References ............................................................................................................. 108
CHAPTER VI............................................................................................................... 111
6. FRACTIONAL LAYER CRYSTALLIZATION AND SWEATING APPLIED TO
ALUMINIUM SCRAP REFINING ............................................................................. 111
Abstract......................................................................................................................... 111
6.1. Introduction ....................................................................................................... 112
6.2 Experimental....................................................................................................... 113
6.3. Results ............................................................................................................... 115
6.3.1. Chemical analyses ...................................................................................... 115
6.3.2. Micrographs ............................................................................................... 117
6.4. Discussion.......................................................................................................... 122
6.5. Conclusions ....................................................................................................... 123
6.6. Acknowledgements ........................................................................................... 124
6.7. References ......................................................................................................... 125
CHAPTER VII............................................................................................................ 127
7. EVALUATION OF THE ALUMINIUM FALLING LIQUID FILM
CRYSTALLIZATION PROCESS STEPS .............................................................. 127
7.1. Introduction ..................................................................................................... 127
7.2. Hydrodynamics of the thin falling liquid films ............................................. 127
7.3. Heat transfer in falling liquid film ................................................................. 129
7.4. The effect of Reynolds number in diffusion boundary layer thickness ...... 130
7.5. Optimization of the aluminium falling liquid film crystallizer ................... 131
7.5.1. Optimal growth rate. .................................................................................. 131
7.5.2. Optimal crystal production rate. ................................................................ 132
7.5.3. Optimal feed temperature........................................................................... 132

7.5.4. Optimum feed rate (Re). ............................................................................. 133


7.5.5. Optimum crystal growth rate for aluminium falling film process. ............. 133
7.6. Design of the aluminium falling film crystallizer ......................................... 134
7.7. Equipment for falling film crystallization process ....................................... 136
7.7.1. Control of the crystallizer........................................................................... 137
7.8. Conclusions and recommendations................................................................ 139
7.9. References ........................................................................................................ 140
SUMMARY ................................................................................................................. 141
SAMENVATTING ...................................................................................................... 145
PRMBLEDHJE.......................................................................................................... 149
ACKNOWLEDGEMENTS ......................................................................................... 153
PUBLICATIONS ......................................................................................................... 155
CURRICULUM VITAE .............................................................................................. 157

CHAPTER I

1. METAL REFINING WITH FRACTIONAL CRYSTALLIZATION:


STATE-OF-THE-ART AND FUTURE PROSPECTS1
Abstract
Historically speaking, three stages can be recognized in the development of fractional
crystallization processes for the purpose of metal refining. A first milestone was the
invention of the Pattinson process for the extraction of silver from lead in 1833, the
second was the development of the zone-melting technique in 1952, and the third one is
the so-called Yunnan crystalliser for refining of tin as of 1975, which is probably the
most advanced application for metals at the moment. Fractional crystallization, widely
known in the metals world as fractional solidification, is a separation technique applied
for purification of metals and organic melts. While most of the currently available
processes are mainly developed for the production of high-purity metals (99.99 wt %
and more), the emerging technologies appear to be exploring the possibilities for
aluminium scrap recycling. For the latter application, proposed methods have not left
the laboratory stage yet with research in this field still ongoing. General restrictions are
that the processes are quite slow and of limited production capacity. Also, the different
requirements on the formation of the crystals (concerning crystal growth rates and
stirring) and the separation of the solid from the liquid fraction are in most cases not
fully met. Consequently, there still is room for further progress. Therefore, a substantial
part of this paper is concentrated on exploring the possible technological solutions to
these challenges. State-of-the-art techniques used in melt crystallization processes for
purifying of organic compounds for polymer production industry and their potential
adoption for metal refining, as well as potential metals that are suitable for the use of
similar techniques are explored.

This chapter was published by Drini B., Katgerman L., and Boom R: Proceedings of the ECI
Conference on Metal Separation Technologies III, June 20-24 2004, Copper Mountain, Colorado, USA,
Eds. Aune, R.E. and Keukkonen M., Helsinki University of Technology, Laboratory of Metallurgy,
Espoo, Finland; p. 34-41.

Chapter I

1.0. Introduction
The worlds metals production relies more and more on recycling of used metals.
Absolute leader in metal recycling is the gold industry for reasons of scarcity and value
per unit of weight. The steel industry, as biggest producer of metal approaching an
annual production level of 1 billion tonnes of crude steel, is an important recycler with
about 68 % recycled steel.
Aluminium, with 22 million tonnes production of row aluminium in 2003 the second
metal in terms of quantity, is increasing rapidly the fraction of recycled (secondary)
aluminium. The different aluminium alloy series, however, are not easily compatible in
terms of mixing: a beneficial element in one series is harmful for other series. Therefore
separation as well as purification of recycled aluminium alloys has received much
scientific and technological attention starting in the eighties of the last century. On top
of that the need for ultra-pure aluminium has increased, mainly for hard disc
applications in personal computers.
Melt crystallization is a separation process by which the fractional separation is effected
by directional crystallization from the melt. Fractional crystallization is applied for both
ultra purification of organic compounds [1, 2], and metals [3, 4]. Historically speaking,
three stages can be recognized in the development of fractional crystallization processes
for the purpose of refining. A first milestone was set by the invention of the Pattinson
process [5] for the extraction of silver from lead in 1833; further highlights were the
origination of the zone-melting technique in 1952 by Pfann [6], and a particular
elaboration for refining of tin as of 1975 [7]. This last example the so-called Yunnan
crystalliser is probably the most advanced application for metals at the moment. As a
continuous process and in an integrated functional design, the Yunnan crystalliser is
being used industrially (yields of 3045 tonnes/day).
In this paper one particular technique available for purification of metals, fractional
crystallization, is described and compared to the same technique used for purification of
organic compounds for the polymer industry.

Metal refining with fractional crystallization: State-of-the-art and future prospects

Fractional
Crystallization
Suspension
Crystallization
Batch

Layer
Crystallization

Continuous

Static

Dynamic

Figure 1. Overview of fractional crystallization processes.

1.1. Theory
Melt crystallization is a separation technique applied for purification of metals and
organic melts. It is widely known as fractional crystallization in crystallization worlds
and fractional solidification in the metal world.

An overview of fractional

crystallization processes is shown in figure 1. The principle of this technique is to cool


the melt in a controlled way which triggers a slow crystal growth rate. The effect is that,
depending on the phase diagram, crystals will get free of impurities or in any case an
increased purity in one or the other component depending if the system is monotecticeutectic or peritectic.
Most melt crystallization processes use a melt that is being cooled (fractional
solidification), but it is also possible to use a solid that is being heated (fractional
melting). A further distinguishing feature (figure 2) for the crystallization is whether the
crystals grow freely suspended in the melt (suspension-based methods), or as a fixed
layer on a cool surface (layer-based methods). An advantage of the former type is that
the solid/liquid interface is relatively large, which is favourable in terms of crystal
purity and production rate. A strong point of the Layer based process is that it does not
need additional processing to separate the solid from the liquid fraction. Further, the
degree of mixing is an important feature in these processes: when done without, pilingup of impurities in front of the advancing solid-liquid interface may induce unstable
growing of dendrites and, as a result, inclusion of the impurities in the solid fraction. An
overview of the suspension and layer based processes is given in table 1.

Chapter I

crystal

Cooled
wall

melt

Crystal
layer

melt

Figure 2. Suspension (on the left) and layer (on the right) grown crystals.

Metals have a tendency of forming solid solutions. This has an adverse effect on
efficiency of purification of the metal in one crystallization stage. Therefore, if a
successful purification is to be achieved, the crystallization process has to be repeated.
This makes the process less energy efficient and also reduces the yield of the pure
crystal product. Even non solid solution forming systems are not 100 percent pure after
one crystallization stage. This is mainly due to the inclusion of the mother liquor (base
liquid metal) inside the crystal pores and adhesion of impurities on the surface of the
crystals. Therefore a further purification of crystals has to be performed. There are two
main processes applied to organic compounds for this purpose: Sweating and Washing.
Sweating is defined as a temperature induced post-purification step based on a partial
melting of crystals or crystal layers close to the melting point of the pure substance. As
a result, the impurities entrapped in the pores of crystals and adhered to their surface remelt, partially dilute with the pure crystals and drain off under the influence of gravity
forces. The temperature rises continuously, decreasing the viscosity of the entrapped
impurities and easing the draining process. The retention time of a sweating stage
applied to organic melt crystallization varies from 15 minutes for the falling film
process to several hours for static layer growth process.

Metal refining with fractional crystallization: State-of-the-art and future prospects

Table 1. Comparison between layer and suspension growth processes [2].


Parameter

Layer growth

Suspension growth

Crystal surface area

Up to100 m2 m-3

Up to10 000 m2 m-3

Growth rate

10-6-10-5 ms-1

10-8-10-7 ms-1

Mixing intensity

Pure for static process,

Good

Good for dynamic one


Separation efficiency

Moderate

Excellent

Purification of solid

Suitable

Less suitable

Solid liquid separation

No

Yes

Slurry handling

No

Yes

Mode of operation

Repetitive batch

Continuous

Design

Simple

Complex

Scale-up

Multiplication of units

Engineering

Operational attention

Normal

Above average

Maintenance intensity

Normal

Above average

solutions

There are two different washing processes. The first one is diffusion washing, where
washing liquid causes liquid/liquid diffusion of impurities from surface and pores of
crystals to the washing liquid. The second one is rinsing process, where the highly
contaminated adhered impurities at the crystal surface are washed of by pure washing
liquid. The first process has a retention time of typically 15 minutes or longer, while the
second one lasts only a few seconds. For both washing processes the washing liquid
should be overheated in order to prevent its crystallization on the crystal surface.

1.2. Layer based processes


There is a large variety of the layer based processes for fractional crystallization of
metals and the field has been reviewed by several authors [3, 4]. In this paper we will
discuss some of the most important fractional crystallization processes applied for
metals both layer based as well as suspension based.

Chapter I

1.2.1. Heinzer et al. Crystallizer


Heinzer et al. [8] developed a layer based process for purification of silicon. The
crystallizer is shown in figure 3. It consists of steel reactor (1) which has a double
walled outer casing through which cooling water can flow through pipes (2) and (3).
Inside the reactor is a molybdenum sheet beam reflector (4), graphite crucibles (6, 11,
and 31) which are used as respectively melting, crystallization, and residue collecting
vessels. All the vessels are heated by resistance heaters. The operation process consists
of three main steps: i) melting of silicon to be purified in vessel (6), which is added
through pipe (5); ii) slow crystallization of silicon on a rotating cylinder (12) partly
immersed in the liquid and the surface of the cylinder is cooled with argon through a
hollow shaft (13) below the melting point of the liquid silicon and crystals are formed
on its surface; and iii) The upper part of the cylinder is heated up by water cooled
copper coils (25) to facilitate the melting of the crystal layer and the purified product is
collected in a separate container through channel (28). This process is operated in a
continuous way by continuously adding batches of molten silicon to the crystallization
bath and periodical removal of the remaining impurity enriched liquid from the bath.
The inert atmosphere is created by pumping out the air through pipe (9) and flowing in
argon from pipe (8).
This process has been successfully applied for purification of liquid metals such as
silicon and germanium. However, its application for purification of other metals such as
aluminium was not very successful. The main problems encountered are the
impregnation of the interdendritic porosity with mother liquor, oxidation of the layer,
and problems with effective removal of the solidified crystal layer from the rotating
cylinder.

Metal refining with fractional crystallization: State-of-the-art and future prospects

Figure 3. Layer based process for purification of silicon and germanium by fractional
crystallization [8].

1.2.2. Pechiney crystallizer


Figure 4 depicts the aluminium Pechiney crystallization process, which is based on
layer growth on a horizontal rotary drum (2) [9]. The cooled cylinder (2) turns slowly
in the tightly fitting trough (5), consisting of an inlet section (6) and an outlet section
(7), which are separated from each other. The direction of turning of the cylinder is
shown by arrow (3). The material crystallizes in the inlet section and when the layer
passes to the outlet section it is melted under the influence of heating. To provide for
crystallization, the top of the cylinder is cooled with cooling material (10), which could
be liquid or gaseous. The remaining liquid in the inlet zone is enriched with impurities
and is depleted due to the removal of the crystals formed in this zone. The process is
continuous and from time to time the feed material (13) is added to the section (6),
while mother liquor (15 ) and purified liquid metal (14) are removed. The amount of the
feed material added should be equal to the amount of the pure material and mother
liquor moved from the crystallizer.
7

Chapter I

This process is a continuous process for purifying of aluminium. The process can
integrate several stages in one crystallizer where the purified material of one stage can
serve as a feed material for the next stage and so on to desired amount of purification
stages. This can be done by adding additional cylinders parallel to each other (cascade).

Figure 4. Aluminium Pechiney crystallizer [9].

Despite advantages of being a continuous and possibly multistage process, this process
has its disadvantages in that it is very sensitive to the tightness of the trough that
separates the inlet and outlet sections, lack of mixing on the crystallization zone to
provide improved conditions for mass and heat transfer, and difficulties of scaling up.

1.3. Suspension based Processes

1.3.1. Alcoa crystallizer


Alcoa process for purification of aluminium with suspension based fractional
crystallization technique is shown in figure 5 [10]. In this process, molten aluminium is
introduced into container (60) for purification. The molten aluminium is either primary
aluminium (typically 99.6 % purity) or higher purity aluminium produced in an
electrolytic cell known as Hoops cell [11] (99.9 to 99.993 % purity). The heat is

Metal refining with fractional crystallization: State-of-the-art and future prospects

removed from the top of the crystallizer to initiate formation of the crystals in the zone
(70) and under influence of gravity they fall down to zone (72).

Figure 5. Alcoa fractional crystallization process for purification of aluminium [10].

The bottom of the container is heated to prevent formation of incrustation and in the
same time to provide for sweating (partial melting) of the formed crystals, which
increases their purity and yield significantly compared to a previous version of this
crystallizer [11] that did not have this option. After a certain amount of crystals is
formed the crystals are pressed down by the tamper (78), which in addition to
compacting crystals to zone (72), breaks the agglomerates of crystals. The impure
aluminium mother liquor (74), high in eutectic impurities, is removed from the upper
part of the container, through the upper port (76). The remaining pure crystals are
melted and removed through the lower port (80).
While this process achieves significant purification results, it is a batch process and is
not suited for large scale operations that a recycling plant would require. Additionally, it
is a static process, which means that there is no mixing provided to enhance heat and
mass transfer, and encrustation to the walls of the crystallizer are inevitable, leading to
diminished purification efficiency.

Chapter I

1.3.2. Yunnan Crystallizer


The so-called Yunnan process was developed in 1975 [7] and is probably the most
advanced continuous process for refining of metals with fractional crystallization. This
crystallizer is used for refining of tin alloys and has proven itself as successful apparatus
for this purpose in the last 25 years in tin refineries all over the World.
Qiu et al. [7] describe the operational principle of the continuous crystallizer as
presented in Figure 6. The Yunnan Crystallizer consists of electric heaters (3) placed
between inner (1) and outer troughs (2) beneath a spiral agitator consisting of a screw
axes (4) and a spiral blades (5). The spiral agitator is installed in the inner trough of the
crystallizer and is driven by a motor (10). The speed of the agitator is controlled by a
gearbox (11). The angle of crystallizer can be arranged by means of an inclination
regulator fitted to the base (6).
The crystallizer is heated up to the melting point of the feed material and than a feed
material is added in liquid form till the screw axes are immersed fully in the melt and
then the motor is started. Water is sprayed on the surface of the liquid feed material to
perform a controlled crystallization. The crystals are separated from the liquid phase
and transported upwards towards outlet (9), while the impurity rich liquid runs
downward towards outlet (8) under the influence of gravity and inclination of the
crystallizer. This way a counter-current flow of impure liquid and pure crystals is
achieved.
The temperature in the crystallizer is controlled in such a way that temperature of the
lower end of crystallizer is kept close to the eutectic temperature of the alloy to be
refined and the upper end of the crystallizer is heated to the melting temperature of the
purified metal. The adjustment of temperature profile, described above, takes time to
reach steady-state and when it is achieved a continuous amount of feed material is
added to the inner trough of the crystallizer. In the same time a continuous pure material
from outlet (9) and eutectic rich material from outlet (8) are discharged, making a
continuous operation of the crystallizer possible.

10

Metal refining with fractional crystallization: State-of-the-art and future prospects

Figure 6. Yunnan continuous crystallizer: 1- inner trough, 2- outer trough, 3- heaters,


4- screw axis, 5- spiral blade, 6- base, 7- charge pot, 8- outlet for liquid phase, 9- outlet
for crystals, 10- motor, and 11- gearbox [7].
Although the Yunnan crystallizer is a very advanced process for refining low melting
temperature metal alloys such as Sn-Pb and Pb-Ag alloys, its application for other
higher melting temperature alloys of other metals is not possible. The main obstacles to
that are increased heat radiation, and therefore heat loss, when operating at higher
temperatures, which makes temperature control of the crystallizer much more
complicated. Another problem is that most of the metal alloys have a very short
crystallization temperature window, sometimes only a few K, which makes the Yunnan
crystallizer impossible to control under those conditions. A further problem in using the
Yunnan crystallizer for other metals could be that the metallic screw would dissolve in
the molten metal.

1.4. Emerging technologies


Where the currently available processes, described in section 1.2 and 1.3, are mainly
developed for the production of high-purity metals (99.99 wt % and more), the
technology appears still to be under-explored for aluminium scrap recycling. For the
latter application, proposed methods have not left the laboratory stage yet with research
in this field still ongoing. General restrictions are that the processes are quite slow and
of limited production capacity. Also, the different requirements on the formation of the

11

Chapter I

crystals (concerning crystal growth rates and stirring) and the separation of the solid
from the liquid fraction are in most cases not fully met. Consequently, there still is room
for further progress.
A Dutch group, consisting of Corus plc. in IJmuiden, Dutch Organization for Applied
Research (known from its Dutch initials as TNO), and Delft University of Technology
are working on a fractional crystallization process for refining aluminium scrap. While
Delft is working on a layer based process, other partners in Dutch group are working in
a suspension based fractional crystallization process.
Our research group at Delft University of Technology works on a fractional
crystallization process for aluminium recycling based on static layer growth on a
vertical cooled tube with closed bottom, which is shown in figure 7 [12]. The main
emphasis in this process is paid to sweating as a post-purification process [13]. After a
crystal layer of approximately 2-3 cm is grown the cooled tube with an aluminium
crystal layer is removed from the mother liquor and is put on the sweating oven.

Cooling
air

Argon

10 cm

Crystal layer

TC

5 cm

20 cm

Figure 7. Static layer crystallizer for aluminium scrap refining [12].

Sweating is based on the partial melting of the crystal layer by gentle heating close to
the melting point of the pure substance. The impure melt adhered to the surface of the
crystal layer or present in the intercellular regions of the crystal layer (see figure 8)
drains off under the influence of gravity. The temperature during the sweating process is
higher than in the crystallization stage. Although during the sweating process about 5
to10 percent of the crystals are removed as a sweat material, sweating is effective and

12

Metal refining with fractional crystallization: State-of-the-art and future prospects

avoids an additional crystallization step in a fraction of the time (~15 min). Aluminium
with 6 wt. % silicon was used as a model alloy. This alloy was used for reasons of
simplicity of the phase diagram and relatively large temperature span of crystallization.

a)

b)

Figure 8. Optical micrographs for the AlSi6 alloy from layer growth experiments: a)
before sweating; b) after sweating (dark spots represent emptied intercellular eutectic).
The line in the picture denotes 0.1 mm [13].
The static layer crystallization experiments showed that typical growth rates of 1 to 3
m/s are optimal for successful application of sweating as a post-purification process.
The obtained results [12, 13] showed that the same purity of a crystal layer can be
achieved by applying sweating to a layer grown with 2 m/s with the layer grown at
half the speed (1 m/s) without sweating, which means that twice higher production rate
can be applied to achieve the same purification effect (figure 9).

7
6

% Si

5
4
3
2
1
0
starting alloy

1 m/s

2 m/s

2 m/s, after
sweating

Figure 9. Results for AlSi6 alloy from layer growth experiments: before sweating at
growth rate 1 and 2m/s; and after sweating of crystals grown at 2 m/s [13].

13

Chapter I

The suspension process developed by the Dutch group [14] is built in such a way that
stirring of the crystallizer, to enhance mass and heat transfer, is provided with a metallic
stirrer coated with refractory coating material. When approximately 20 to 25 % of the
aluminium is crystallized, crystals are filtered off from the mother liquor. The crystals
are much purer than from the layer based process. However, without an effective
separation process this process can not succeed to overcome problems such as low
crystal yield and removal of adhered impurities on the crystal surface and in crystal
pores.
Alcoa has also adopted the suspension based fractional crystallization process,
described in section 1.3.1, for aluminium scrap refining purposes [15]. Its performance
is in line with expectations, with understandably lower purification efficiency [15, 16],
which is known for processes that have higher alloying element concentration in the
starting material.

Figure 10. A schematic illustration of the Japanese suspension crystallizer for refining
aluminium scrap [17].

The most advanced group on research in aluminium scrap recycling is the Japanese
group [17]. They developed a suspension based process with capacity of producing
about 1000 tonne per month of purified aluminium scrap. The crystallizer is shown in
figure 10. It is similar to other processes described above with one distinction that a
compression filter is applied from the top and the crystals are packed in the bottom of
the crystallizer.

14

Metal refining with fractional crystallization: State-of-the-art and future prospects

Zn%

Si %
1
0.8
0.6
0.4
0.2
0

Mn%

Fe%

Cu%
Initial concentration
Refined composition

Figure 11. Refining results of the radiator aluminium scrap with the Japanese
suspension crystallizer [17].

In figure 11 results of the purification efficiency of the radiator scrap are shown. It is
clearly seen that the removal efficiency of alloying elements is much lower than it is
possible from binary alloy phase diagrams. The reason for that is the presence of iron
and manganese as alloying components which form intermetalic compounds and thus
reduce the efficiency of their removal from aluminium scrap.
The main problem with suspension based processes is that there is no efficient process
of separating of the formed crystals from the mother liquor. Therefore, if a sustainable
suspension crystallization process is to succeed a better way of separating of crystals
from mother liquor has to be developed.

1.4. A look to the future


Fractional crystallization technology is in a much more advanced stage for organic
materials than for metals. This is mainly due to the ease of handling organic materials,
as their melting points are lower than for metals and other factors such as liquid
pumping, slurry handling, corrosion and costs of equipment are more favourable for
them. On the other hand, metals have their advantages concerning low viscosity, high
heat conductivity, higher density and diffusivity, which have a direct positive effect on
production and purification rate. The heat of fusion is comparable for both groups per

15

Chapter I

unit of weight, but it differs a factor of 8 per unit of volume. The latent of fusion has to
be removed on solidification and is therefore a liability for metals. However, due to the
much higher thermal conductivity for metals this liability is compensated [6].
In table 2 different parameters for organic materials and metals that can be purified by
the fractional crystallization technique are given. The data show that the main problems
for adopting technologies that are applied for organic compounds are high equipment
costs, which require special materials that can resist to the high corrosive attack of most
of the liquid metals, and difficulties of handling metal slurries. Other problems for
liquid metals could be their oxidation in contact with air, and high operational
temperatures required. The later one could pose serious problems concerning equipment
corrosion and energy efficiency. However, solutions for these problems are available in
a metallurgical process industry.

Table 2. Comparison of different parameters between organic compounds and metals


that are purified by the fractional crystallization technique.
Parameter

Metals

Organic compounds

Melting temperature (Tm) K

300 - 1800

60 % are 273 -473

Liquid diffusivity (D) m2 s-1

10-10

10-9

Liquid viscosity () N s m-1

low

Could be a problem

20 - 400

0.1 - 0.6

200 - 400

200 - 380

0.1 1.2

1.3 3.3

1700 - 10500

<1000

Slurry handling

Difficult

Easy

Equipment corrosion

High

Low to moderate

Ease of pumping

Difficult

Easy

Equipment costs

High

Medium

Thermal conductivity (kth) W m-1 K-1


Latent heat of fusion (Hf) kJ kg
Specific heat (Cp) kJ kg-1 K-1
Density () kg m

-3

-1

For suspension-based refining, the use of so-called washing columns, originally


developed for organic chemicals, could be a promising option in order to solve the
problem of impure liquid adhering to the crystal surface during the solid/liquid
separation. In these, the crystals are rinsed counter-currently in a flow of a pure melt.
The combination of a separate crystalliser (in which a semi-solid slurry is produced)
with a washing column (in which the slurry is split in a product and a residue) could

16

Metal refining with fractional crystallization: State-of-the-art and future prospects

then be a conceivable solution [18]. Except for the improved purification, benefits of
this concept would be that it is easily (up)-scalable with a high specific capacity.
Further, it is a continuous process, which has an advantage over batch processes
concerning energy efficiency. However, in order for this technique to succeed new
pumps should be invented that are capable of pumping metal slurries without heating it.
Corus recently patented a process for purification of metals with continuous suspension
based fractional crystallization [19, 20], which is shown schematically in figure 12. The
crystallization apparatus (1) consists of a chamber (2) in which a layer of a cooling
liquid (4) and a layer of a partially molten metal (5) to be purified are present. The
cooling liquid is drown in (arrow A) and circulated through pipe (6), where a cooling
device (8) is present, by means of pump (7) and after cooling is turned back to chamber
(2) (arrow B). The density of the cooling liquid is higher than that of the molten metal
and they should not mix with each other to avoid contamination. The molten metal is
added through inlet (10) and in contact with cooling liquid crystallization takes place.
The crystals remain for a while in suspension as a result of mixing, than fall on the top
of the cooling liquid and are transported by it towards outlet (11). In the same time the
molten metal enriched with alloying elements move counter currently towards outlet
(12). The chamber is divided in several compartments and a temperature gradient is
imposed as a result of heating up of the cooling liquid due to the contact with hotter
molten metal and dissipation of the energy of crystallization of the molten metal. The
process is continuous with the amount of the added molten metal in inlet (10) equalling
the amount of pure crystal removed in outlet (11) and mother liquor removed from
outlet (12). The same process is possible with the cooling liquid having lower density
than molten metal to accommodate removal of the crystals of metals which have lower
density of crystals than their liquid [20].

17

Chapter I

Figure 12. Schematic presentation of a cross section of the Corus crystallizer [19].

Main advantage of the Corus crystallizer is that it is a continuous process; it provides


cooling from a liquid salt and avoids formation of incrustations on the cooling wall,
which is the main obstacle to suspension based processes; and as a result of using salts
as cooling liquid an accurate temperature control can be achieved. This is the most
promising advanced suspension based method for purifying metals by employing
fractional crystallization technique.
In layer-based refining, two operating modes are possible: static or dynamic. In the
static mode, mass transfer of impurities away from the growing crystal surface occurs
by diffusion only and low growth rates of 10-710-6 m/s are required (for organic
compounds) in order to attain the required purity. In the dynamic mode, mass transfer is
enhanced by turbulence allowing growth rates typically one order of magnitude higher
than in the static case. Falling films seem to be the best option for generating turbulence
at the crystal surface, as has been demonstrated on an industrial scale for a range of
organic chemicals [1, 2]. In a falling-film crystalliser [21], the melt is circulated from a
feed vessel over a vertical heat exchanger as a thin film over the wall. On each pass, a
small fraction of the melt crystallises on the wall, while the remaining melt is recirculated until a predefined amount of the feed material has solidified. The layer grown
on the heat exchanger is the actual purified alloy and is recovered by re-melting. Purity
of the layer can be enhanced further by applying one of the post-purification processes:
sweating, diffusion washing, or rinsing. The purification efficiency of the sweating
process for static layer crystallization of aluminium [13] is comparable to one
18

Metal refining with fractional crystallization: State-of-the-art and future prospects

crystallization stage. Results from organic compound purification processes show that
their purifying effect for falling film [1] is even greater than for static processes.
Fractional crystallization for metals is widely used for lead, tin, and silver, and could be
applied to purify technologically important metals such as silicon, aluminium and
magnesium. Liquid silicon is a metal, and the large-scale production of solar cells based
upon silicon demands an economic purification technology that produces ultra-pure
silicon. The yield of a solar cell is strongly dependent on silicon purity. For aluminium,
there is a demand for ultra-pure aluminium (made from primary metal) for application
in the electronics industry (hard disc memories). There is also a fast growing need to
purify secondary aluminium from recycled scrap. Magnesium, as a rapidly growing
metal with applications in light-weighting design for transport and in casings for lap top
computers and mobile phones, will have to be recycled and purified because tramp
elements dissolved in magnesium strongly deteriorate the corrosion resistance of this
metal.
The future of fractional crystallization for these and possibly other metals will be bright
due to the economic, energy conservation and ecological drives.

1.5. Conclusions
Fractional crystallization technique is being used successfully for metal refining for
almost two centuries. The most advanced process for this purpose is the Yunnan
crystallizer for separating tin from lead. While existing technologies are mostly of a
limited production capacity and are mainly used to produce super-pure metals, the
increasing demand for developing this technique for recycling of metal scrap requires
processes that have much higher production capacity. Therefore research in this field is
intensifying rapidly.
Fractional crystallization technique is much more advanced for organic compounds than
for metals. This is mainly due to the better process conditions, such as low melting
temperature and lower reactor attack, of organic materials. However, a careful study of
possibilities of adopting these processes for metals will open new horizons. Prospects
are good for both modes of operation: suspension and layer crystallization. The most
important metals that could use this technique are aluminium, silicon, magnesium, lead,
and precious metals.

19

Chapter I

1.6. Acknowledgements
This research is carried out as part of the EET funded project Sustainable Technology
for the Reclamation of High-grade Aluminium from Scrap (contract number
EETK98026). Fruitful discussions, about the organic material technologies, with Prof.
P.J. Jansens of the Laboratory for Process Equipment of Delft University of Technology
are greatly appreciated.

1.7. References

1. Arkenbout, G.J.: Melt Crystallization Technology. Lancaster PA USA:


Technomic Publishing Company Inc., 1995.
2. Jansens, P.J. and Matsuoka, M.: Melt Crystallization, in: Encyclopaedia of
Separation Science, Eds. Wilson I.D., Adlard E.R., Cook M., and Poole C.F.,
Academic Press, San Diego, New York London, Sidney, Tokyo, (2000) pp. 966975.
3. Sillekens, W.H., Verdoes, D., and Schade van Westrum, J.A.F.M.: "Refining
Aluminium Scrap by means of Fractional Crystallization: Technical Feasibility",
Proceedings of the Fourth ASM International Conference and Exhibition on the
Recycling of Metals, ASM Europe, (1999) pp. 105-114.
4.

Qiu, K. Sudholter, S., Kruger, J., and Yang, X.: Raffination von Metallen
durch fraktionierte Kristallisation aus der Schmelze [Historical development
and prospects: metal refining by fractional crystallization from the melt], Metall.
Vol. 49, no. 7-8 (1995) pp. 491-495.

5. Pattinson: Separating Silver from Lead, British Patent no.6497, 23 November


1833.
6. Pfann, W.G.: Zone Melting. John Wiley & Sons Inc., New York, NY, U.S.A.,
2nd edition,1966.
7. Qiu, K., Duan, W., and Chen, Q.: Basic Principles of Control of Continuous
Crystallizer in Metal Refining, Transactions of the Institution of Mining and
Metallurgy, Section C. Vol. 110 (2001) pp. C161-C164.

20

Metal refining with fractional crystallization: State-of-the-art and future prospects

8. Heinzer, H., Rath, H-J., and Schmidt, D.: Process for Purifying Solid
Substances, US Patent no. 4231755, Application: 19 May, 1978, Accepted: 04
November, 1980.
9. Boutin, R.F., Process for the Continuous Purification of Metals by Fractional
Crystallization on a Rotary Drum, US patent no.4581062, Application: 13 May,
1985, Accepted: 8 April, 1986.
10. Dawless R.K. and Graziano R.E. Fractional Crystallization Process US Patent
no. 4294612, Application: Nov. 30, 1979, Accepted: Oct 13, 1981.
11. Dawless R.K. and Jacobbs S.C., Production of Extreme Purity Aluminium,
US4273627, Application: Dec 26, 1979, Accepted Jun 16, 1981.
12. Mehmetaj, B., Bruinsma, O.S.L., Kool, W.H., Jansens, P.J., and Katgerman, L.:
Aluminium Scrap Recycling with Solid Layer Fractional Crystallization,
Proceedings of the 15th International Symposium on Industrial Crystallization
(ISIC-15),

Ed.

Chianese,

A.,

Sorrento,

Italy,

Chemical

Engineering

Transactions, Volume 1, 2002, pp.879 884.


13. Mehmetaj, B., Kool, W.H., Jansens, P.J., and Katgerman, L.: Fractional Layer
Crystallization and Sweating Applied to Aluminium Scrap Recycling,
Proceedings of the 7th International Conference on Semi-solid Processing of
Alloys and Compounds (7th S2P), Eds. Tsutsui Y., Kiuchi M., and Ichikawa K.,
National Institute of Advanced Sci. & Techn.- Japan Society for Technology of
Plasticity, Tsukuba, Japan, 24-28 September 2002, pp. 611 - 616.
14. Nienoord, M., Verdoes, D., Boender, W., Landskroon, J., Sillekens, W.H.:
Batch Crystallization Experiments for the Refining of Aluminium Scrap,
Proceedings of the 15th International Symposium on Industrial Crystallization
(ISIC-15),

Ed.

Chianese,

A.,

Sorrento,

Italy,

Chemical

Engineering

Transactions, Volume 1, 2002, pp.879 884.


15. Kahveci, A.I. and Unal, A.: "Refining of a 5XXX Series Aluminum Alloy Scrap
by Alcoa Fractional Crystallization Process", Proceedings of the Fourth
International Symposium on Recycling of Metals and Engineered Materials, Ed.
D.L. Stewart Jr., J.C. Daley, and R.L. Stephens, Warrendale PA, USA: TMS,
22-25 October 2000, pp. 979-991.
16. Dawless R.K., Troup R.L., Meier D.L., and Rohatgi A., Production of Extreme
Purity Aluminium and Silicon by Fractional Crystallization Processing, J. of
Cryst. Growth, 89 (1988) pp. 68-74.
21

Chapter I

17. Sotome T. and Ohtaki M., "Application of Fractional Crystallization for


Refining of Molten Aluminium Scrap", Conference Proceedings of ICAA 6, vol.
1, Ed. T. Sato et al. Tokyo, Japan: The Japan Institute of Light Metals, (1998)
pp. 351-356.
18. Verdoes, D., Visscher, H.: Method and Device for Separating Metals and/or
Metal alloys of Different Melting Point, World patent no.1998/27240,
Application: 09 December 1997, Accepted: 27 Jun 1998.
19. De Vries, P.A., and Wouters, H.A.: Method for Fractional Crystallization of a
Molten Metal, World patent no.2004/005558, Application: 27 Jun 2003,
Accepted: 15 January 2004.
20. De Vries, P.A., and Wouters, H.A.: Method for Fractional Crystallization of a
Metal, World patent no.2004/005559, Application: 27 Jun 2003, Accepted: 15
January 2004.
21. Drini, B., Kool, W.H., Jansens, P.J., and Katgerman, L.: Method of Recovering
a Metal from a Mixture, PCT/NL2004/000202, Filled for international patent
on 25 March 2004.

22

CHAPTER II
2. SCOPE

2.1. Fractional crystallization principle and aluminium refining


When cooling a crystalline material from the melt, the lattice structure of the crystals as
they begin to grow cannot easily accommodate secondary constituents, so that these are
accumulating in the remaining liquid while the solidified fraction is relatively clean. This
is the basic mechanism behind refining by means of fractional crystallization. A plain
example presented by nature is that of pack ice: formed from seawater by freezing, but
comprising of freshwater. As such, the treatment is not new, and is being applied on an
industrial scale for the (ultra)-purification of organic chemicals as well as for metals.
Refining by means of fractional crystallization requires the presence of a semi-solid (i.e.,
solid/liquid) state, bordered by a eutectic or a monotectic equilibrium. The principle at
hand can be illustrated on the basis of figure 2.1, in which the phase diagram for a simple
binary system is sketched.

Figure 2.1. Binary Alloy Phase Diagram for a Eutectic System: Schematic Enlargement of the Semi-solid
State (+L).

Chapter 2

Table 2.1. Limits on the Purification of Secondary Constituents from Aluminium by means
of Fractional Crystallization (Single-step Operation)
Element

System

Maximum Temperature

Thermodynamic Distribution

Range T (C)

Coefficient k0 ()

Si

eutectic

83

0.13

Fe

eutectic

Cu

eutectic

Mn

eutectic

Mg

eutectic

210

Cr

peritectic

Ni

eutectic

21

0.00

Zn

eutectic

279

0.87*

Ga

eutectic

633

0.20*

peritectic

Zr

peritectic

Ti

peritectic

Pb

monotectic

1.6

0.00

Sn

eutectic

432

0.00*

Cd

monotectic

11

0.05

Sb

eutectic

3.5

0.00

peritectic

Bi

monotectic

3.5

0.04

5.5

0.03

112

0.17

2.5

0.62
0.45*

* By approximation
In evaluating the performance of a fractional crystallization process, an indicator called
the distribution coefficient k is used. Relating to a mixture of the components A (here:
aluminium) and B (any other element), it is defined as:

26

Scope

k=

CB,s
.
CB,u

(1)

CB,s and CB,u are the concentrations of component B in the solidified fraction and in the
untreated material respectively. This distribution coefficient, sometimes referred to as the
effective distribution coefficient, depends on the thermodynamic properties of the system,
but also on the process kinetics. The distribution coefficient in the equilibrium condition
can be derived from the phase diagram of the concerned system and is called the
thermodynamic distribution coefficient k0.
A central question in considering the use of fractional crystallization for refining
aluminium is to what extent secondary constituents can be theoretically purged from the
base metal by this principle. Particularly, the constraints on the purification and yield are of
interest, as well as the requirements concerning such parameters as processing temperature.
An answer on this can be obtained from the phase diagrams of the concerned systems [1].
Restricting to binary alloy diagrams here, mixtures of aluminium and a single other
element theoretical limits are summarised in table 2.1.
These are grouped from top to bottom as major alloying elements, minor alloying elements
and tramp elements in aluminium.
Considerations in appraising this information are recapitulated next:
Assumption is that both fractions in the semi-solid state are separated from each other
completely; that is, without mutual attachment. Thus, the obtained distribution coefficient
constitutes a lower limit. The refining performance does essentially depend on the content
of the alloying element in the feed material, but this shows to be of minor relevance for
most of the considered binary systems. For those elements, where it is of significance,
approximations (averages) are given. Listed values are obtained by using weight
concentrations.
The refining performance does also depend on the yield P of the process, which is defined
as the quantity or flow of the product as compared to that of the unrefined input (0P1). In
the table, only the data for P=0 (no yield!) are represented.
Relating to the matter of temperature control for such processes, the maximum temperature

27

Chapter 2

range T is included in the table 2.1. It is taken as the full temperature interval of the semisolid state (i.e., TmATE in figure 2.1). Thus, it is an upper limit that may be much smaller,
depending on the actual alloying contents.
Most of the listed systems are either eutectic or monotectic. They are thus apt for refining
by means of fractional crystallization processes in which the impurities are concentrated in
the remaining melt. Some systems are peritectic, meaning that the concerned secondary
constituents end up in the opposite fraction (that is, the crystals). It can be argued that these
elements can be removed by some other treatment, like filtering the crystals from the melt.
For complicated alloy systems (like actual scrap aluminium), however, it remains
problematic that a material that is purified with respect to one category of constituents is
enriched with the other. For the major alloying elements all being eutectic the
theoretical prospects are good for iron, silicon, and copper, fair for magnesium, and
moderate for manganese. Apart from this, process control will for some cases be critical:
the available temperature interval is very small for iron and manganese.
Finally, a reservation that has to be made is that the conditions during the actual refining
operation will be principally different from the quasi-static conditions that apply for the
phase diagrams. Further, it should be noted that an extrapolation to systems with more than
two components is only indicative.

2.2. Selection of alloy systems to be studied


Within a joint project of Delft University of Technology with Corus plc in IJmuiden,
TNO-MEP in Apeldoorn and TNO Institute of Industrial Technology in Eindhoven,
carried out as a part of the research project Sustainable Technology for the Reclamation
of High-grade Aluminium from Scrap, which is undertaken in the context of the Dutch
R&D stimulation programme on Economy, Ecology and Technology (contract number
EETK98026), research partners worked on the modification of two fractional
crystallization processes so that can be applied to the purification of aluminium scrap.
For practical reasons, only a few alloys can be studied. Therefore, only a few alloy
systems and a few concentrations of the alloying elements can be selected. The alloy
systems that are investigated in this thesis have been chosen in two ways. These two
28

Scope

approaches are: a selection based on the wishes of Corus Group's aluminium business
units, and a selection based on generally relevant binary and ternary phase diagrams.
Binary phase systems are suitable to practice using the "tools". Ternary phase systems,
e.g. Al-Si-Fe and Al-Si-Mn, are the first systems that are relevant to industrial
applications of fractional crystallization.
Concentrations of alloying elements in the inlet stream of each alloy system have been
selected on the basis of guesses at the compositions of scrap types. The following four
systems have thus been selected :

Al - Si ;

Al - Si - Fe ;

Al - Si - Mn ; and

Al - Cu - Mg.

In the Al-Si system, there is an eutectic at about 12 mass % Si. To have a good starting
point for the process during the initial experiments, an alloy halfway between pure Al and
the eutectic, an Al-Si alloy with 6 mass % Si, has been chosen. This system is only used
to hone skills.
A ternary system that is of interest is the Al-Si-Fe system. Both Si and Fe are always
present in aluminium alloys. Two sources of Fe in aluminium scrap are foil, and parts that
have been produced with high pressure die casting[1,2]. Liquid aluminium reduces SiO2,
which comes from sand particles or refractory material. This makes that Si will be present
in molten scrap. Hence, aluminium scrap contains both Fe and Si, and the Fe content may
have to be reduced anyway. Therefore, it has been suggested to study the ternary Al-SiFe system in the project on fractional crystallization. In general, the alloys for pressure
die casting contain at least 0.6 mass % Fe to reduce sticking of the cast part to the die [1].
Food packaging is the main application for aluminium foil. An alloy that is being used for
foil production is AA8014 [3]. It contains less than 0.3 mass % Si and 1.2 to 1.6 mass %
Fe. Therefore, an alloy with 0.3 mass % Si and 1.2 mass % Fe has been selected.

29

Chapter 2

The third alloy system studied in this project is the Al - Si - Mn system, more specifically
an alloy of composition 1.5 mass % Si and 1.0 mass % Mn, because this is a typical
mixed scrap composition of braizing sheet.
The fourth system is the ternary system Al-Cu-Mg with 4 mass % Cu and 1.5 mass %
Mg. This composition represents more or less the composition of chips generated during
the machining of sheets or plates of AA2XXX alloys that are used in aircraft structures
[1,3]. These chips are not used as "raw material" for the production of wrought alloys at
the moment as they form a mixture that contains too many alloys. An avenue to close the
recycle loop for AA2XXX alloys may be the application of a purification technique like
fractional crystallization.

2.3. Research goals


The main research goal of this thesis is to prove fractional layer crystallization process, in
a laboratory scale, as a tool for purification of different aluminium scrap streams. Such a
process should operate well within all relevant alloy systems, and it should be able to
purify sufficiently the aluminium scrap types that are of interest to industry. Thus, the
chance of unpleasant surprises during the use of a fractional crystallization process in a
real plant is reduced.
Within the main research goal several subgoals can be distinguished, such as:

Designing and constructing of the laboratory scale static crystallizer for


aluminium refining.

Studying the optimum layer growth rate for most economical way of purifying
selected aluminium alloys.

Examininning the structure of the grown crystal layers and impurity distribution
within crystal layers.

Testing the efect of post-purification techniques, such as sweating, on overal


purification of crystal mass.

Designing and construction of a falling film crystallizer for aluminium refining.

30

Scope

All the selected aluminium alloys were studied with zone melting, static layer growth,
and sweating postpurification technique. Suitable layers were obtained and studied with
different analysing techniques to unlish the impurity distribution and their local
concentration, the size of crystal cells and intercelular spaces depanding on growth rate,
etc. The obtained layers were also sweated to remove the impurities that are concentrated
mainly in the intercellular space of crystal layers and the effect on overall purification
was established.
The possibility of purifying aluminium scrap with falling film fractional crystallization
technique was tested theoreticaly and equipment for this purpose was constructed.

2.4. Thesis outline


Chapters of this thesis are writen in such a way that they can be read indipendentaly and
most of them are published or sent for publications to journals and conference
proceedings.
In chapter 1, a review of fractional crystallization process and its application for refining
different metals and organic materials are described. A detailed theoretical principles of
fractional crystallization, including different models for zone melting and fractional
crystallization processes are described in chapter 3. Chapter 4 deals with studying the
feasibility of fractional crystallization process through a simple experimental technique,
such as zone melting. Layer based fractional crystallization techniqu for refining selected
binary and ternary aluminium alloys is presented in chapter 5, while the use of
postpurification technique called sweating is described in chapter 6. Finaly, Chapter 7
describes theoretical possibilities and equipment design of falling film fractional
crystallization process for aluminium refining.

31

Chapter 2

2.5.

References

1. J.R. Davies et al., Metals Handbook, Tenth Edition, Volume 2 Properties and
Selection : Nonferrous Alloys and Special-Purpose Materials, ASM International
(1990)
2. D.G. Altenpohl, Aluminium: Technology, Applications, and Environment, The
Minerals, Metals & Materials Society, Warrendale, PA, USA (1998)
3. J.P. Lyle and D.A. Granger, "Aluminum Alloys", Ullmann's Encyclopaedia of
Industrial Chemistry, 5th Edition, Volume A1, Ed. W. Gerhartz, VCH
Verlagsgesellschaft mbH, Weinheim, Germany (1985)

32

CHAPTER III
3. SOLUTE REDISTRIBUTION DURING DIRECTIONAL
CRYSTALLIZATION AND INTERFACE MORPHOLOGY

3.1. Introduction
As a result of the differences in solute solubility in liquid and solid state, there is a
solute redistribution at the solid liquid interface. If the solubility of the solute in the
crystal phase is smaller than in the liquid phase, the solute is rejected at the interface
leading to solute pile up in the adjoined liquid, while if the solute solubility in the
crystal phase is higher than in the liquid the solute layer adjoined to the interface is
depleted in solute. The fact that solidus and liquidus lines of the phase diagrams do not
coincide, except for pure materials and in the exceptional cases of congruent melting,
indicates that a solid usually differs in composition from the liquid in equilibrium with
it. Therefore, crystallization processes are always accompanied with solute
redistribution which is different from the homogeneous distribution in the starting
liquid, although the total amount of solute does not change.
Throughout this chapter, except for equilibrium condition, diffusion of the solute in the
solid is not taken into account. The reason is that diffusion in liquid state is much faster
than diffusion in the solid state. In the case of metals, diffusivities in the solid state at
the melting point are by a factor of 100 to 1000 lower than in the liquid state. This is
expected, as diffusion is defined as a transport of mass from one region to another on an
atomic scale. Therefore there is much lower resistance to mass transfer in the liquid than
in the crystal phase.
In this chapter the solute redistribution for a flat solid/liquid interface is described in
sections 3.2 and 3.3. In section 3.4 the morphology of the interface is treated. Cellular
structure of crystals and solute redistribution in such morphology are presented in
sections 3.5. In section 3.6 some selected models used to describe experimental solute
redistribution are analyzed. Finally, general conclusions are presented in section 3.7.

Chapter III

3.2. Solute redistribution during crystallization under equilibrium


conditions
An important parameter to describe solute redistribution phenomena is the
thermodynamic distribution coefficient. The equilibrium distribution coefficient k0 is
defined as the ratio of the concentration of solute in the solid (CS) to that in the liquid
(CL) when the solid and liquid phases are at equilibrium.

k0 =

CS
CL

(3.1)

It is noted that several names are in use for k0. It is variously called as partition
coefficient, segregation coefficient, and distribution coefficient. The last term is the

Temperature

Temperature

most widely used.

CS

CL

Composition
a) k0 < 1

CL
CS
Composition
b) k0 >1

Figure 3.1. Schematic phase diagrams showing solidus and liquidus lines; a) k0 < 1; b)
k0 > 1.
In figure 3.1 is clearly seen that solidus and liquidus are represented with straight lines,
which is a simplification and can be true only for concentration spam. For a negative
slope, k0 is smaller than one, which means that the solute reduces the melting point of
the pure solvent or metal and crystallization leads to purer crystals than the original melt
composition (figure 3.1a). Cooling of a liquid of composition CL, of which the phase
diagram has a positive slope (figure 3.1b), results in formation of crystals with
composition CS which are enriched with impurities.
During equilibrium crystallization the concentration of solute in crystals changes
towards the overall concentration as a result of diffusion in the solid. Figure 3.2

34

Solute redistribution during directional crystallization and interface morphology

schematically shows a phase diagram for which k0 < 1. When a liquid of composition
CL is cooled to a temperature T1 the first crystals will appear with composition CS1. As
crystallization proceeds towards temperature T2, the impurity concentration in the liquid
is represented by point M. The amount of formed crystals is given by the lever rule as
LM/KM. The composition of crystals is uniform and is given by point K (CS2). When
crystallization continues till the entire melt is solidified, the solute concentration of the
crystals is given by point O, and is equal to the starting concentration CL. Therefore,
true equilibrium crystallization gives a uniform concentration which is equal to that of
the starting liquid C0.
This is not often achieved due to the limited diffusion in solid or liquid. In many cases it
also not desired. In case of purification processes a concentration difference between
crystals and liquid is even imperative.

Temperature

T1
T2

T3

CS1 CS2 CL
Composition
Figure 3.2. Phase diagram and solute distribution during equilibrium crystallization.

The solute concentration in the solid part, CS, of an ingot undergoing equilibrium
crystallization, after a fraction fS is crystallized is given by equation 3.2.

CS =

k 0C0
1 + f S (k 0 1)

(3.2)

35

Chapter III

CS, C0, and k0 are expressed in mol, atom, or weight fractions, which are independent of
density.

3.3. Solute redistribution while no diffusion in the solid


To determine the composition profile of the solute at the solid liquid interface the
following conditions will be considered under the assumption that no diffusion in the
solid takes place:

Complete mixing in the liquid

Stagnant liquid, and

Partially mixed liquid.

3.3.1

Complete mixing in the liquid

Complete mixing of the liquid will never be attained since there will always be a liquid
layer in contact with the growing solid interface where solute transport is diffusion
controlled. However, when the diffusion coefficient is high or the growth rate is low,
the condition of complete mixing will be approached.
Under the condition of complete mixing, the solute concentration profile of the crystals,
crystallized directionally from a cylindrical charge, shows a profile which is defined by
the solidus line. Figure 3.3 gives the solute concentration profile for a certain fraction
crystallized. The concentration profile shows a discontinuity at the solid/liquid interface.
The ratio of the solid to liquid concentration after a fraction x is crystallized is given by
CS(x)/CL(x), which is equal to k0 (equation 3.1).
The concentration profile as a result of the crystallization is given by Scheils law [1]
and reads

C S = k 0 C 0 (1 f S ) ( k0 1)

(3.3)

36

Solute redistribution during directional crystallization and interface morphology

Composition
CS(0) = k0C0

CL(x)

C0
CS(x)

SOLID
0

LIQUID

x
Fraction crystallized

Figure 3.3. Solute redistribution during directional crystallization in case of complete

liquid mixing; k0 < 1.


where CS is the solute concentration at the crystal interface after fraction fS has
crystallized.

3.3.2. Stagnant liquid


In case of a stagnant liquid, the only solute transfer mechanism near the interface is by
diffusion. The bulk of the melt retains its original concentration C0. Tiller et al. [2]
treated this case assuming a flat interface, a constant growth rate, constant value of k0
and negligible convection in the liquid.
Based on these assumptions, a solute boundary layer will develop and the bulk of the
melt will retain its original concentration C0, see figure 3.4. During the initial stage of
crystallization the solute concentration at the boundary layer CL(0) increases from C0 till
C0/k0, and the concentration in the solid increases from k0C0 till C0. This means that the
effective distribution coefficient (which will be defined in section 3.3.3.) is equal to one
(keff = 1). When the solidification is nearly completed, the solute boundary layer will
give rise to an enriched final transient layer.

37

Chapter III

CS(0) = k0C0

Composition

Initial
transient

C0
k0
C0

CL(0)

Steady state

Final
transient

CL(x)

CS
0

1
Fraction crystallized

Figure 3.4. Solute redistribution in the solid, CS, and solute concentration in the liquid

at the solid-liquid interface, CL(0). The dashed lines, CL(x) indicate the solute
concentration in the liquid ahead of the solid interface at different fractions
crystallized; k0 < 1 [2].
Tiller et al. [1] derived the following relationship between the position x in the liquid in
front of the crystal interface and the solute concentration in x, CL(x), during steady
state growth:
Gx'
C L ( x' ) = [C L (0) C 0 ]exp
+ C0
D

(3.4)

where G and D are growth rate and diffusion coefficient in the liquid respectively, and
x is distance from the solid interface. The graphic representation of equation 3.4 is
depicted in figure 3.5.

38

Solute redistribution during directional crystallization and interface morphology

Composition

CL(0)
CL(x)

C0

Interface
0
Distance from interface, x

Figure 3.5. Solute redistribution ahead of the solid-liquid interface during directional

crystallization when steady state is achieved; k0 < 1 [2].


Equation 3.4 represents an exponential decay of the solute concentration in the liquid
adjacent to the advancing crystal interface, of which the characteristic distance is
defined by the ratio D/G. Since D is assumed constant, the effect of the growth rate, G,
is the controlling factor determining the decay curve of the solute in the liquid adjacent
to the interface.
Under the assumption that in the initial transient the increase of CS is also exponential
and by applying boundary conditions, Tiller et al. [2] derived the equation for the solute
concentration in the solid, CS(x), at any point x in the crystal in the initial transient:

Gx

C S ( x) = C 0 (1 k 0 ) 1 exp k 0
+ k 0
D

(3.5)

Equation 3.5 describes an asymptotic rise of the solute concentration from k0C0 at x = 0,
to C0, when steady state begins (see figure 4). The length of the initial transient, XI, is

XI =

D
Gk 0

(3.6)

Equation 3.6 shows that XI increases for lower values of k0 and G. When steady state is
achieved, CS = C0 and CL(0) = (C0/k0). If we substitute these expressions in equation 4
the solute concentration in the liquid at the interface at steady state condition is given by

39

Chapter III

1 k0
Gx'
C L ( x' ) = C 0 1 +
exp

k0
D

(3.7)

Steady state will be maintained as long as there is a sufficient amount of liquid ahead of
the crystal interface for forward diffusion of the solute and G remains constant. When
the amount of liquid becomes insufficient, the solute concentration in the crystals begins
to increase above C0 in order to accommodate the excess solute, which was accumulated
in front of the interface. This stage of crystallization is called the final transient. Its
length is D/G and the final transient occupies a shorter distance than the initial transient
by a factor k0 (see figure 3.4). Characteristic values for D/G are 0.01-0.1 cm. The length
of the initial transient is dependent on the alloying system. During directional
crystallization when the liquid is stagnant, only the initial transient region becomes
purified.

3.3.3. Partial mixing in the liquid


If mixing in the liquid is not complete but only partial and the crystal growth rate does
not proceed slowly, solute atoms are rejected by the advancing interface at a faster rate
than they can diffuse to the bulk of the liquid. Hence, similar to the case of the stagnant
liquid, a concentration gradient is created in front of the advancing crystals. Figure 6
shows this effect.
Burton, Prim, and Slichter (BPS) [3] have analyzed the idealized Chochralski
configuration for crystal growth from diluted melts mixed under laminar flow
conditions.
The BPS model assumes that the crystal interface is flat, which means that there is no
radial segregation of the solute. They also assume that the solute transport in a layer
near the interface with the length is diffusion limited (which layer is called the
stagnant boundary layer), and that the solute is completely mixed under the influence of
convection in the bulk of the liquid (see figure 3.6). The last assumption makes the BPS
model different from the model of Tiller et al. [2].

40

Composition
CS(0) = k0C0

Solute redistribution during directional crystallization and interface morphology

CL(x)
CL

C0
CS(x)

SOLID

LIQUID

x
Fraction crystallized

Figure 3.6. Solute redistribution during directional crystallization in case of partial

liquid mixing; k0 < 1.


Although it is an approximation to regard the liquid as stagnant over the length and in
motion after it, it is a useful assumption in solving practical problems of fractional
crystallization. The BPS model gives an expression for the thickness of the diffusion
layer for small growth rates. The value of depends on the velocity of the liquid
parallel to the crystallization interface, , on the solute diffusion in the liquid, and on
the kinematic viscosity of the liquid . The expression reads

= 1 .6

1
3

1
6

(3.8)

1
2

Thickness of the boundary layer strongly depends on the intensity of stirring in the
liquid, and ranges from 0.01 mm for strong stirring to 1 mm for natural convection.

The effective distribution coefficient, keff, is now defined as the ratio of the
concentration of the solute in the solid at the interface, CS(x) (see figure 3.6), to the
concentration of the solute in the bulk of the liquid CL. It should be mentioned that keff
is considered constant over the crystallization process.

41

Chapter III

The value of keff is higher than the value of k0, and ranges from k0, for complete mixing
in the liquid, to one, for diffusion limited transport such as for the stagnant liquid (see
Section 3.3.2). The value of keff depends on the amount of convection in the melt.
The stagnant film model introduced by BPS establishes a relation between the effective
distribution coefficient and the dimensionless parameter G/D, which is used as the
normalized growth velocity. They obtained the following expression for the effective
distribution coefficient keff during steady state growth:

k eff =

C S ( x)
=
CL

k0
k 0 + (1 k 0 )e

(3.9)

G
D

Equation 3.9 can be arranged to


1

G
1
ln
1 = ln 1
k

D
k0
eff

for k0<1

(3.10a)

1
ln1
k
eff

for k0>1

(3.10b)

= ln1 1
k

G

D

In this way, the /D value can be obtained experimentally by determining keff measured
from crystals grown at different growth rates under identical stirring condition and
plotting the values of ln(1/keff 1) or ln(11/keff) against the crystal growth rate. This
should give a straight line with slope -/D. The parameter keff is generally determined
from a crystallization experiment by quenching and measuring the solute concentration
in the solid near the interface and in the bulk liquid.

42

CS(0) = k0C0

Composition

Solute redistribution during directional crystallization and interface morphology

CL(0)
CL(x)
C0

C0

CL

CS
0

1
Fraction crystallized

Figure 3.7. Solute redistribution in the solid (CS), solute redistribution in the liquid

(CL), and solute concentration in the liquid at the solid-liquid interface [CL(0)]. The
dashed lines, CL(x) indicate the solute concentration in the liquid ahead of the solid
interface at different fractions crystallized; k0 < 1.
Figure 3.7 shows that the solute concentration in the solid increases as a result of
increased liquid concentration, which happens due to the rejection of the solute from the
crystal interface.

3.3.4. Concluding remarks


The solute distributions in the crystals after uniaxial directional crystallization,
discussed so far, are summarized in figure 3.8 in case of complete crystallization. It is
clearly seen that the solute redistribution in case of equilibrium crystallization (curve a),
has no effect on the final solute redistribution in the crystallized sample. In case of a
stagnant liquid condition (curve b) only during the short initial transient purification
takes place, whereas the rest of the ingot retains its original concentration. The
impurities removed in the initial transient will be collected at the final transient. For
partial mixing in the liquid (BPS-curve c) and for complete liquid mixing (Scheil-curve
d), the bulk of the ingot gets purified and the last part of the ingot gets enriched with
solute. The separation effect is highest for the Scheil condition.

43

CS(0) = k0C0

Composition

Chapter III

b
a

C0

c
d

1
Relative location in the crystal

Figure 3.8. Solute redistribution in the crystal after complete crystallization, for k0 <

1, in case of: a) complete diffusion in solid and liquid (equilibrium), b) stagnant liquid,
c) partial mixing in liquid (BPS), and d) complete liquid mixing (Scheil).

3.4. Crystal - Melt interface morphology

3.4.1. Constitutional supercooling theory


It is well known that the liquid of an advancing liquid-solid interface of an alloy with a
distribution coefficient smaller than one will have a higher impurity concentration than
the liquid in the bulk of the melt. As a consequence of this, the liquidus temperature of
the liquid close to the interface of the solidifying alloy is lower than the liquidus
temperature of the bulk of the liquid. Hence, supercooling can occur if the temperature
of the interface is below the actual temperature of the bulk of the liquid (crystallization
is achieved by the heat flow from liquid to solid). This phenomenon is known as
constitutional supercooling. The variation of the interface temperature and
concentration for a simple eutectic binary alloy is shown in figure 3.9. It shows that
although the liquidus temperature (Tl) close to the interface is smaller than the liquidus
temperature of the bulk liquid, it is higher than the actual temperature (Tq) imposed by
the heat flow during crystallization, and hence the liquid close to the interface is
constitutionally supercooled due to the increased solute concentration resulting from the
rejection of the solute by the crystal interface.

44

Solute redistribution during directional crystallization and interface morphology

As a result of constitutional supercooling the local liquidus temperature, Tl, of the liquid
will be affected. Tl is related to the composition by
Tl (C0) - Tl = m(C0 Cl)

(3.11)

where Tl (C0) is the liquidus temperature corresponding to the initial alloy composition,
m is the liquidus slope in the phase diagram, C0 is the initial concentration of the alloy,
Cl is the solute concentration ahead of the advancing interface, and Tl is the liquidus
temperature ahead of the advancing interface. Liquidus temperatures determined from
equation 3.11 are indicated in figure 3.9.

C0/k0

C0

Cl
C0
x
Tl

Tl (C0)
Tq

T0
Ts (C0)
C0

Tcool

Tl
C0/k0

Figure 3.9 Variation of liquidus temperature and solute concentration ahead of an

advancing interface during layer crystallization of a simple eutectic binary alloy [1] .

For eutectic systems the liquidus temperature of the liquid ahead of the solidifying
interface increases with increasing distance, x, from the interface. In order to determine
the stability of the growing interface, it is important to determine the temperature
imposed by the heat flow during crystallization, Tq. The temperature gradient in the
liquid, GT,l, is defined as

45

Chapter III

dTq

GT ,l =
dx x =0

(3.12)

Depending on the value of GT,l in the liquid at the solid-liquid interface, there may, or
may not exist a zone of constitutional supercooling (see figure 3.10).

Tq

Tl

Tl

Tq

dTl dTq

dx'
dx'

dTl dTq
>
dx'
dx'

T
Tcool

heat
exchanger wall
Figure 3.10. Condition for constitutional supercooling at the crystal layer interface,

and the resulting interface morphology [5].

A quantitative model of predicting the occurrence of constitutional supercooling was


developed by Rutter and Chalmers [4]. They predicted that cell formation could be
prevented by reducing the impurity content, reducing the speed of growth or increasing
the value of GT,l, which eliminates the region of supercooling (see figure 3.10). The
critical ratio of the temperature gradient and the growth rate GT,l/G is that at which the
length of a supercooled zone is zero or the slopes at the interface of liquidus
temperature and actual temperature lines are equal (figure 3.10 left). The condition for
which constitutional supercooling is present is

GT ,l
G

<

mC 0 1 k 0
*
D
k0

(3.13)

46

Solute redistribution during directional crystallization and interface morphology

where G is the growth rate, D is the diffusion coefficient, and k0 is the thermodynamic
distribution coefficient.
Experimental work followed to prove the constitutional supercooling hypothesis [5-8].
It was conducted by separating solid and liquid by decanting and inspecting the crystal
interface for signs of instability. The results were presented as a GT,l/G versus C0 plot,
which represents a straight line passing through the origin. In this way a distinction was
possible between a stable and unstable interface. The slope was compared with (-m)*(1k0)/(k0D) (see equation 3.13) and a reasonable value of D was found to comprise a
substantiation of the constitutional supercooling theory.
These experiments gave a semiquantitative merit to the constitutional supercooling
theory. However, one must notice that a precise quantitative test was beyond the scope
of these experiments. The quantities k0, m and especially D are not well known, while
GT,l is difficult to measure accurately during the crystal growth from the melt.
The constitutional supercooling principle is very useful and easy to apply even to
complex situations. Hurle [9] has given the expression that explains the condition under
which a stirred melt can be constitutionally supercooled. It reads

GT ,l
G

<

mC0 1 k0
*
*
D
k0

k0
k0 + (1 k0 )e

G
D

(3.14)

The factor, added at the right-hand side of equation 3.14, as compared with equation
3.13, is the same as in the BPS expression for keff and is always smaller or equal to 1 for
k0<1. This represents the fact that stirring enhances the interface stability.
3.4.2. Interface stability theory
The constitutional supercooling theory provides a useful estimation of the solid/liquid
interface stability in directional crystallization conditions. However, theoretically a
drawback of the constitutional supercooling criterion is that it takes into account only
the thermodynamic properties of the liquid, while ignoring the thermodynamic
properties of the crystal and the interface surface tension. Mullins and Sekerka [10-13]
studied the latter properties in their theory of interface stability.

47

Chapter III

Growth direction
0

1
2
3

Figure 3.11. Perturbation forming at the solid/liquid interface. Perturbation grows in

the overall growth direction and expands laterally with wavelength 0 [13].

In planar crystal growth the interface is considered to be initially planar and then a small
perturbation is imagined to appear (see figure 3.11). Whether the perturbation will grow
or disappear depends on the surface energy, on the solute and thermal fields, and on the
interface kinetics. Mullins and Sekerka [11, 12] made the assumption that there is
equilibrium at the solid/liquid interface, that no convection is present, and that the
surface energy is isotropic. The instability criterion is given by

GT ,l
G

mC 0 1 k 0 s + l
H
*
*
*S
<
D
k0
2 l
2 l

(3.15)

where H is the heat of fusion, s and l are the thermal conductivities in solid and
liquid respectively, S is a dimensionless stability function, and the other parameters are
the same as in equation 13. The S function is a function of k0 and is graphically given as
a function of the dimensionless parameter A in figure 12. A is defined by

A=

k 02
G TM
*
*
1 k 0 D mC 0

(3.16)

where is the solid/liquid surface energy and TM is the melting temperature of the pure
solvent. Parameter A varies linearly with and its value is 10-3 or smaller.

48

Solute redistribution during directional crystallization and interface morphology

The value of S is typically 0.8 to 0.9. For negligible small solid/liquid surface energy
S=1. Hence, the deviation of S from 1 represents the tendency of the surface free energy

to stabilize the interface. After neglecting the term H/2kl, equation 3.15 can also be
presented as [13]:

__

GT
S
GC m

(3.17)

where GC is solute concentration gradient at the interface and GT is the conductivity


averaged mean temperature gradient.
Apart from the deviation of S from 1 due to the surface tension, the instability criterion
(equation 3.15) varies from the constitutional supercooling one (equation 3.13) in, that it
uses GT (see equation 3.17) instead of the temperature gradient in the liquid. GT is
defined by

GT =

GT , s s + GT ,l l

(3.18)

s +l

where GT,s is temperature gradient in the solid at the interface.


The value of GT is usually higher than that of the temperature gradient in the liquid
(GT,l). For the case of aluminium for example, s / l = 2.3, while for water it is around
4. This means that the difference between the interface stability criterion and the
constitutional supercooling one is more pronounced for water than for aluminium.

49

Chapter III

k0=0.01
k0=0.1
k0=0.3
k0=0.5
k0=0.9
k0=5

0.9
0.8

k0=1000

0.7

k0=100

k0=10

0.6
0.5

Unstable

0.4
0.3
0.2

Stable

0.1
0
-10

-9

-8

-7

-6

-5

-4

-3

-2

-1

log10A

Figure 3.13. Plot of the dimensionless stability function S versus log10A for different k0

values [13].

It can be seen from equation 3.15 that the latent heat of fusion H contributes to the
stabilizing term H/2 l , which can have an important effect in diluted solutions. Its
stabilizing effect is small for metals, but can have a dominant impact for instance for the
ice/water system [14].

3.5. Cell structure and solute redistribution in cellular solidification


As explained in the previous section, with a limited amount of constitutional
supercooling, the interface morphology of the growing crystal will change from planar
to cellular. Cells have a crystal morphology which usually grows anti-parallel to the
heat flow direction. If a higher deviation from the constitutional supercooling principle
is present, the cell interface will get branched and dendritic growth will take place. The
cellular crystal morphology is the least wanted growth form, if an effective solute
50

Solute redistribution during directional crystallization and interface morphology

redistribution from directional growth is to be achieved. It will be further treated in this


section.

a)

b)

Figure 3.13. Morphology of the growth front of an Al-Cu alloy: a) Initial perturbation

of the planar front at grain boundaries; b) Transverse section of the fully developed
cells [14, 15].

Sharp and Hellawell [14, 15] studied the interface morphology changes of an Al-Cu 2
percent alloy by a sudden acceleration of the growth rate and subsequent quenching of
the sample to observe the morphological changes in the crystal interface. They found
that the breakdown of the flat interface starts close to the grain boundaries, which act as
built-in distortions of the planar interface. This effect is clearly seen in figure 3.13a. The
interface develops further to a fully cellular structure as shown in figure 3.13b, which
shows that cells are regularly spaced.
A useful way to observe the cellular structure as it develops is by using transparent
organic compounds. Succinonitrile is the most widely used organic substance for this
purpose. The main reason for that are its plastic crystal properties, and thus, like metals,
it crystallizes with a macroscopically smooth solid/liquid interface and forms no facets.
This behavior reflects the isotropic properties of metals and a special class of molecular
crystals. A typical succinonitrile cellular interface morphology is given in figure 3.14a.

51

Chapter III

a)

b)

Figure 3.14. Morphology of the growth fronts of transparent organic crystals; a)

Succinonitrile cellular front [16]; b) Salol growing with faceted cells [17].

The cellular morphology of the transparent organic crystals of which the solid/liquid
interface develops facets is shown in figure 3.14b. Most of the non metallic crystals,
including semiconductors, develop facets due to their anisotropic crystal nature. The
conditions of formation of faceted cells are similar to the one described above for
metals.
During cellular growth the solute concentrations in the cells and in the intercellular
space differ considerably. This observation has been confirmed by numerous
researchers [18-20]. They found that the degree of segregation towards the intercellular
space can be so high that a lead rich phase was present even in a tin 0.16 % lead alloy.

52

Solute redistribution during directional crystallization and interface morphology

Figure

3.15.

Concentration

profile

ahead

of

the

cellular

interface

for

succinonitrile0.075wt% acetone, Cf/k indicates the tip concentration, k is the


thermodynamic distribution coefficient [21].

Assuming a simple binary alloy like succinonitrile acetone [21] containing a eutectic
point, the maximum solute concentration in the intercellular region is equal to the
eutectic concentration (CE = C0/k0). This condition is depicted in figure 3.15.
In sections 3.2 and 3.3, the solute redistribution was considered only for directional
crystallization with a flat solid/liquid interface. Due to this assumption the equations
describing the redistribution are relatively simple. These equations are not valid to
describe the cellular growth solute redistribution. The multidimensional diffusion alone
makes it impossible to describe the redistribution in a satisfactory way. Additionally,
there are complications of solid state diffusion and radius of curvature.
Some attempts were made for describing the solute redistribution in nonplanar growth
conditions [22-27]. These microsegregation models are semi-analytical and approximate
relationships. They simplify the conditions by treating only a volume element which has
a small thickness in the growth direction and extends towards the neighboring cell wall
[22]. These are simplified models which cannot describe accurately all the processes
that go on during cellular crystal growth. Their results are in line with the experimentaly
achieved ones, but their accuracy is compromised by making numerous approximations.

53

Chapter III

3.6. Selected models used to describe experimentally achieved solute


redistribution in directional crystallization

3.6.1. Cheng, Irving and Kyle model


Modifications to the BPS model are suggested by several authors. Cheng et al. [28]
who analyzed the crystallization of benzene-hexadecane and used equation 3.10a. They
found out that there were two regions of different linearity over the range of growth
rates measured, see figure 3.16. After analysis of the interface morphology, they found
out that each region is associated with a distinguished crystal interface. The first region
of linearity at growth rates up to 1.2 cm per hour corresponded with a cellular structure,
while for higher growth rates the interface was dendritic. Trapping of the solute in the
intercellular or interdendritic regions at the interface seems the main reason for such
anomalous behavior of this alloying system.
When trapping occurs in this eutectic benzene-hexadecane system the average solute
_

solid concentration at the interface, C S , is given in terms of solute concentration in the


liquid at the interface and the void fraction of the irregular solid/liquid interface f.

C S = fC L (x)

(3.19)

Equation 3.19 is equivalent to equation 3.1 with the surface void fraction corresponding
to the equilibrium distribution coefficient. Further, Chang et al. [28] gave a modified
expression for the effective distribution coefficient given by BPS in equation 9, where
instead of the equilibrium distribution coefficient k0, a surface void fraction f parameter
is given.

k eff =

f
f + (1 f )e

G
D

(3.20)

54

Solute redistribution during directional crystallization and interface morphology

1.2
1
1
ln
1
k
0.8
eff

0.6
0.4
0.2
0
0

Growth Rate

10
cm/hour

Figure 3.16. Logarithmic 1/(keff-1) versus growth rate for the organic system

hexadecane in benzene, using equation 10a [28].

As mentioned above, the two regions of linearity for the alloying system correspond
quite closely to the two distinct regions of interface morphology. This brings us to the
conclusion that a constant or average value of interface void fraction f can be given to
each of the regions of linearity. The value of f is higher than that of k0.

3.6.2. Impurity inclusions in crystal layer


As discussed in the previous section, during the layer crystallization process certain
amount of impurities are incorporated in the purified crystal mass. In this section the
reasons of their incorporation and distribution in the crystal layer will be described. The
mechanism of impurity concentration, the operational measures to prevent impurity
incorporation, and the post purification measures are presented in figure 3.17 [29, 30].
The layer, crystallized in a crystallization experiment, consists of three (sub)layers, see
figure 3.16. In the first layer, close to the heat exchanger, in the nucleation stage
impurity inclusion is high due to the rapid crystallization to initiate the crystallization
process. The impurities are also concentrated in the second layer of the crystal layer as a
result of departure from equilibrium conditions of crystallization. The amount of
inclusions in the layer is increased with increasing growth rate. Finally, impurities are
concentrated in the surface of the crystal layer as a result of adherence of the
contaminated residual melt at the end of the crystallization process. By far the largest

55

Chapter III

portion of crystals is in the second layer and the nucleation and adherence layers are
very thin. The post purification steps mentioned in figure 16 are described in the next
section.
Cooled

melt

Wall)
Step

(1)

(2)

(3)

Mechanism of
impurity
incorporation
High impurity
concentration
due to high
nucleation rate at
high supercooling

Impurities on
crystal layer due
to high crystal
growth rates

Impurities due to
adherence of
contaminated
residual melt at
the end of the
process

Operation measures to
prevent impurities in the
crystal product
-Nucleation by controlled
temperature and flow
conditions
-Controlled seeding
-Mechanically induced
nucleation by shock waves,
ultrasonic vibration, or
cavitations
Purity criteria which relates
the maximum allowable
growth rate for the
production of pure layer

Additional
purifying
measures
-sweating

Minimized and smooth


crystal layer surface to
reduce the wettable area

-sweating
-washing

-sweating

Figure 3.17. The mechanism of impurity concentration, operational measures to

prevent impurity incorporation, and post purification measures [29, 30].

Modeling of the layer crystallization processes is empirical [31], mainly due to the
complexity of various kinetic effects described in previous sections. During the
crystallization process the impurity segregation effect is measured by the effective
distribution coefficient (keff), which accounts for equilibrium, kinetic and phase
separation effects. Therefore, the measurement and use of the keff values must be used at
constant conditions with respect to growth rate, initial impurity concentration, etc.. In
practice the main advantage of the model described below is that it gives an expression
of the purification achievable in a single stage depending on the amount of the original
liquid remaining un-crystallized. Now we consider an initial mass of liquid m0, with
impurity concentration C0, being partially crystallized to give a reduced mass of liquid
mr, with an impurity concentration Cr. Since keff can be considered as a constant, the

56

Solute redistribution during directional crystallization and interface morphology

mass and component balance for a differential balance of crystallized mass can be
integrated to give an average distribution coefficient kcr:

k cr = 1 +

log(C r / C 0 )
log(mr / m0 )

(3.21)

0.4
k_cr

0.3
0.2
0.1
0
0

0.05

0.1

0.15

0.2

%PTZ

Figure 3.18. The kcr variation as a function of the initial impurity concentration for the

system Phenothiazine (PTZ) in acrylic acid [31].

Although the average distribution coefficient kcr is often approximated to be a constant


over a given purity range, substantial changes in kcr appear, depending on the
concentration of the impurities in the system. figure 3.18 shows how the kcr varies as a
function of the initial impurity concentration for the system Phenothiazine (PTZ) in
acrylic acid. It is clear from figure 3.18 that separation is less effective at higher initial
impurity concentrations. This can be attributed mainly to the quantitative increase of the
impurity inclusions due to the increased impurity level in the melt which results in a
crystal layer with more entrapped and adhered liquid.

3.6.3. Post purification steps


There are two main post purification steps that are applied in the layer crystallization
process: sweating and washing. More background reading on postpurification processes
can be found in Arkenbout [32].

57

Chapter III

Feed
liquid

One purification
stage
Phase 1:
crystallization

Phase 2:
Partial
melting/

Phase 3:
Total melting

Liquid
Product

Residue
liquid
Cooling

Gentle heating

Heating

Figure 3.19. Schematic description of the layer crystallization process.

Sweating is defined as a temperature induced purification step, which is based on the


partial melting of the crystals by gentle heating close to the melting point of the pure
substance [33]. The impurities adhered to the surface of the crystal layer as well as those
present in the pores of the crystal melt, partially dilute with the pure material and drain
off under the influence of gravity. The temperature rise decreases the viscosity of the
enclosed impurities and eases the process of draining off. Additionally, due to the raised
temperature of the included melt, which is rich in impurities, part of the pure crystals is
dissolved in the impure liquid, opening channels for flow of the impurities out of the
crystal layer. Sweating is effective in removing impurities from all the layers of the
crystal (see figure 3.17). A schematic description of the sweating stage in the layer
crystallization process is described in figure 3.19.
During the sweating process about 10 % of the product is lost [34]. However, the
purification effect of the sweating stage is in the same range of one additional
crystallization step and only a fraction of the time is needed. The sweating time can last
from 15 minutes for the falling film process [34] to several hours for some static
processes of crystallization [35] of organic compounds.

58

Solute redistribution during directional crystallization and interface morphology

The theoretical description of the sweating process is studied by Wangnick [36] for
layer based processes. The proposed equation gives the correlation for predicting the
sweating effect on the additional purification that it can achieve. The equation reads
k eff , sw = 0.028 0.277 * sw + 1.235 * 2sw

(3.22)

k eff , sw = k eff ,before k eff ,after

(3.22a)

with:

sw = k eff ,before *

and

Tsw
Teq

(3.22b)

where keff,sw is the difference between the distribution coefficients before (keff,before) and
after sweating (keff,after) of the grown crystal layer, sw is a characteristic dimensionless
number of the sweating process, Tsw is the temperature of sweating, and Teq is the
equilibrium melting temperature of the crystal.
Equation 3.22 was successfully applied to both static and dynamic layer crystallization
processes for organic compounds [36].
There are two different washing processes applied in layer crystallization processes:
rinsing and diffusion washing [36]. Rinsing is applied to remove the highly
contaminated melt that is attached to the surface of the crystal by purer liquid. The
rinsing process lasts only a few seconds. Diffusion washing causes liquid-to-liquid
diffusion of impurities from the pores of the crystal layer. To prevent the solidification
of the washing liquid onto the crystal the washing liquid is superheated. Therefore
diffusion washing is always accompanied by sweating, which further increases the
purification effect. The duration of the diffusion washing process usually is more than
15 minutes. Diffusion washing is applied to remove the impurities from the surface of
the crystal layer and in the pores of the grown layer (see figure 3.17). As a washing
liquid for the diffusion washing process one can use the starting material, which is purer
than the impurities included in the pores and the surface of the crystal. The washing
liquid can be added to the new starting material for crystallization. The purification
effect of the washing process is lower than for other purification steps. However, it is

59

Chapter III

well suited for easy removal of the impurities that are included in the outer part of the
crystal layer.

3.7. General conclusions


The solute redistribution in case of equilibrium crystallization has no effect on the final
solute redistribution in the crystallized sample, as it assumes complete solute diffusivity
in both solid and liquid phase. In case of a stagnant liquid condition only during the
short initial transient purification takes place, whereas the rest of the ingot retains its
original concentration and the impurities removed in the initial transient will be
collected at the final transient. For partial mixing in the liquid (BPS) and for complete
liquid mixing (Scheil), the bulk of the ingot gets purified and the last part of the ingot
gets enriched with solute. The separation effect is highest for the Scheil condition. All
of these models assume a flat interface. While equilibrium condition is not achievable in
practice the stagnant liquid model is only true for rapid solidification. The BPS is valid
for diluted solutions, provided slow growth rates are applied. The Scheil condition
shows the maximum theoretically possible solute redistribution.
As a result of the phenomenon known as constitutional supercooling, the interface
morphology departs from the flat crystal front and the interface gets rippled. Regarding
the degree of departure from the constitutional supercooling criteria, cellular or
dendritic crystal growth will take place. This makes the description of solute
redistribution phenomena much more complicated and no model is at present available
that can account for all the factors influencing it. The cellular crystal structure is the
most common structure in real systems of directional crystallization.
The practical models describing solute redistribution achieved experimentally are
mostly empirical. They describe the morphology as cellular crystals with impurity
inclusions between them. There are several methods applied to increase the purification
effect of directionally crystallized materials by applying post purification steps such as
sweating and washing.

60

Solute redistribution during directional crystallization and interface morphology

3.8. References
1. W. Kurz and D.J. Fisher, Fundamentals of solidification, Trans Tech
Publications, Uetikon-Zrich, 1992.
2. W.A. Tiller, K.A. Jackson, J.W. Rutter, and B. Chalmers, Acta Met., 1 (1953)
428.
3. J.A. Burton, R.C. Prim and W.P. Slichter, The Journal of Chemical Physics, 21
(1953) 1987.
4. J.W. Rutter and B. Chalmers, Can. J. Phys., 31 (1953) 15.
5. W.A. Tiller and B. Chalmers, Can. J. Phys., 34 (1956) 96.
6. D. Walton, W.A. Tiller, and J.W. Rutter, Trans. AIME, 203 (1955) 1023.
7. T.S. Plaskett and W.C Winegard, Can. J. Phys., 37 (1959) 1555.
8. G.S. Cole and W.C. Winegard, J. Inst. Met., 93 (1964) 322.
9. D.T.J. Hurle, Solid-State Electron., 3 (1961) 37.
10. W.W. Mullins and R.F. Sekerka, J. Appl. Phys., 34 (1963) 323.
11. W.W. Mullins and R.F. Sekerka, J. Appl. Phys., 35 (1964) 444.
12. R.F. Sekerka, J. Appl. Phys., 36 (1965) 264.
13. R.F. Sekerka, J. Crystal Growth, 3-4 (1968) 71.
14. R.M. Sharp and A. Hellawell, J. Crystal Growth, 6 (1970) 253.
15. R.M. Sharp and A. Hellawell, J. Crystal Growth, 6 (1970) 334.
16. A.Pocheau and M. Georgelin, J. Crystal Growth, 250 (2003) 100.
17. K.A. Jackson and J.D. Hunt, Acta Met., 13 (1963) 1212.
18. J.J. Kramer, G.F. Boling, and W.A. Tiller, Trans. AIME, 227 (1963) 364.
19. H. Biloni, G.F. Boling, and G.S. Cole, Trans. AIME, 233 (1965) 251.
20. H. Biloni, G.F. Boling, and H.A. Domian, Trans. AIME, 233 (1965) 1926.
21. B. Kauerauf, G. Zimmermann, S. Rex, M. Mathes, and F. Grote, J. Crystal
Growth, 223 (2001) 265.
22. H.D. Brody, M.C. Flemings, Trans. Met. Soc. AIME, 236 (1966) 615.
23. T.W. Clyne, W. Kurz, Metallurgical Transactions, 12A (1981) 965.
24. I. Ohnaka, Transactions ISIJ, 26 (1986) 1045.
25. C.Y. Wang, C. Beckermann, Material Science and Engineering, 171 (1993) 199.
26. H. Yoo, C.-J. Kim, Int. J. Heat Mass Transfer, 41 (1998) 4379.
27. V.R. Voller, Int. J. Heat Mass Transfer, 43 (2000) 2047.
28. C. Chang, D.Irving, and B. Kyle, AICHE-Journal, 13 (1967) 739.
61

Chapter III

29. R. Scholz, K. Wangnik, and J. Ulrich, J. Phys. D: Appl. Phys., 26 1993) B156.
30. R. Scholz, Ph.D. Thesis, University of Bremen, Germany, VDI-FortschrittsBerichte, No. 347, VDI-Verlag, Dsseldorf, Germany, 1993.
31. M. Stepanski and E. Schfer, In Couirs on:Theory and Application of Melt
Crystallization, 28th 30th September 2000, Martin Luther University HalleWitwnbwrg, Halle (Saale), Germany.
32. G.F. Arkenbout, Melt Crystallization Technology, Technomic Publishing
Company Inc., Lancaster, Basel, 1995.
33. J. Ulrich and J. Bierwirth, In Science and Technology of Crystal Growth, (J.P.
van der Erden and O.S.L. Bruinsma, eds.) Kluwer Academic Publ., Dordrecht,
The Netherlands, 1995, 245.
34. S.J. Jancic, In Industrial Crystallization 87, (J. Nnyvlt and S. Zacek, eds.),
Elsevier Science Publ., Amsterdam, The Netherlands, 1887, 57.
35. S. Rittner and R. Steiner, Chem.-Ing.-Techn., 57-2 (1985), 91.
36. K. Wangnik, Ph.D. Thesis, University of Bremen, Germany, VDI-FortschrittsBerichte, No. 355, VDI-Verlag, Dsseldorf, Germany, 1994.

62

CHAPTER IV
4. ZONE MELTING INVESTIGATION OF THE ALUMINIUM
SCRAP REFINING POTENTIAL IN FRACTIONAL
CRYSTALLIZATION1
Abstract
Recycling of aluminium is an effective way of enhancing the efficient use of energy and
limiting the burden on the environment. Purification based on fractional crystallization
appears to be a promising option for this purpose. This implies that the material is to be
processed in the semi-solid state; that is, at a temperature where a solid and a liquid
phase co-exist. In this chapter, it is shown that zone melting can serve as a perfect
indicative test for assessing the refining potential of fractional crystallization processes,
since they are both based on the same thermodynamic principle. The results of zonemelting experiments are then presented for some aluminium model alloys representative
of the scrap streams. Both chemical analyses and microscopic examination revealed that
the significant purification that takes place in it extends to more than the half of the
sample. The BPS- (developed by Burton, Prim and Slichter) and Pfan models are used
to interpret these results.

Key words: aluminium recycling, zone melting, fractional crystallization, micrographs,


line scan.

Part of this work was published by Sillekens W.H., Schade van Westrum J.A.F.M., Bruinsma O.S.L.,
Mehmetaj B., Nienoord M.; Proceedings of the Fourth International Symposium on Recycling of Metals
and Engineered Materials, Eds. Stewart Jr. D.L., Daley J.C., and Stephens R.L., TMS, Warrendale PA,
USA (2000): p. 963977.

Chapter IV

4.1. Introduction

Production of aluminium from bauxite is an energy intensive process. For production of


1 kg aluminium as much as 14 kWh of electrical energy is spent, whereas for re-melting
of aluminium scrap a fraction of that energy, only 5-10 percent, is needed. Given these
facts, it is clear that aluminium recycling leads to significant cost savings.
For most applications, however, aluminium alloys need to comply with certain
specifications regarding the contents in alloying elements, as well as with limitations on
contaminating elements. Scrap batches originating from used products do not usually
meet the requirements for wrought alloys, which are generally more stringent than for
casting alloys. Such batches should: either be diluted with primary aluminium and then
re-alloyed up to the specification of the alloy envisaged, or they should be used for
casting alloys (cascade recycling).
With an overall growth of aluminium usage, cascade recycling and dilution with
primary aluminium do not pose immediate problems. In the long run, however, a shift
towards recycling process where aluminium scrap can be used for other than its original
purpose will be needed. Reclamation processes that are capable of providing a constant
high quality of secondary aluminium will be a necessity to achieve this, implying that
research in developing a sustainable aluminium recycling process is now an important
topic [1-4].
Melt crystallization is a separation process by which the fractional separation is effected
by directional crystallization from the melt. Fractional crystallization is applied for both
ultra-purification of organic compounds [5, 6] and metals [1]. Historically, the first
application of fractional crystallization for metals was the invention of the Pattinson
process for extraction of silver from lead in 1829 [7]. It was followed by the
development of the zone melting technique in the 1950s by Pfann [8,9].
In this chapter we have applied a zone refining (known also as a zone melting)
technique to assess the purification potential of the fractional crystallization process in
the context of the aluminium scrap purification potential. Zone refining is a simple
technique of coupled melting and freezing and is used for purifying semiconductors,
organic and inorganic compounds, as well as metals. It was first developed by Pfann
[8,9] to purify germanium for use in semiconductors.

64

Zone melting investigation of the aluminium scrap refining potential in fractional crystallization

This chapter focuses on the issue of refining aluminium scrap; that is, on treating scrap
of unfavourable chemical composition to obtain a batch containing lower levels of
alloying and contaminating elements. Zone melting is introduced as a relatively simple
technique to experimentally assess the refining potential and to predict that in fractional
crystallization. Experimental results for some aluminium model alloys will be presented
and compared with the analytical model developed by Burton, Prim and Slichter [10].

4.2. Theoretical Framework


Refining by means of fractional crystallization requires the presence of a semi-solid
(i.e., solid/liquid) state, accompanied by eutectic or monotectic equilibrium. The
principle for a eutectic system can be seen in figure 1 for the aluminium-rich part (in
this case, component A is aluminium). For a holding temperature Tr, the phase diagram
describes that an alloy of initial concentration CB,u consists of a solid fraction and a
liquid fraction, for which the contents in B are CB,s and CB,l respectively. Separating
both fractions means that an essentially cleaner product (namely the solid fraction) and
an enriched residue (the remaining liquid) are obtained. Refining of the peritectic
systems is also possible with fractional crystallization, however, in this case the pure
material is collected in liquid phase.
In evaluating the performance of a fractional crystallization process, the distribution
coefficient, k, is used. With respect to a mixture of the components A (here: aluminium)
and B (any other element), it is defined as

k=

Cs
Cu

(4.1)

where Cs and Cu are the concentrations of impurities in the solidified fraction and in the
original material respectively. This distribution coefficient, sometimes referred to as the
effective distribution coefficient, depends on the thermodynamic properties of the
system, but also on the process kinetics. Some sources prefer the use of the so-called
removal efficiency, , which is the complementary fraction of the distribution
coefficient k (=1k). This distribution coefficient in the equilibrium condition can be

65

Chapter IV

derived from the phase diagram of the system and is called the thermodynamic
distribution coefficient, k0.

Figure 4.1. Binary phase diagram for a eutectic system.


Burton, Prim and Slichter analysed the process kinetics of fractional crystallization for
the particular case of a rotating crystal withdrawn from a melt [10]. They obtained the
following general relation for the distribution coefficient:

G s
1 1
1 = 1 exp .
k k0
k d l

(4.2)

In this so-called BPS-model, G denotes the crystal growth rate and kd the mass transfer
rate. The latter represents the transfer of (accumulated) impurities away from the
growing crystal, resulting from diffusion, convection or some type of deliberate stirring.
The density of the solid and liquid are designated s and l respectively. The model
indicates that a low growth rate and a high mass transfer rate will result in a high crystal
purity.
A further indicator for the removal of secondary constituents from the primary
component is the purification ratio, R, which is defined as:
66

Zone melting investigation of the aluminium scrap refining potential in fractional crystallization

R=

Cs
Cl

(4.3)

Here, Cl represents the concentration of impurities in the remaining melt or liquid


fraction. In reality, it is a measure of the extent to which the impurities accumulate in
the residue.
During the zone melting of a sample there are to competing processes: in one side
crystallization of the material, while on the other side melting of crystals takes place
simultaneously. Pfann [9] describes the changing concentration of the crystals for a
single pass of the melting zone as given in equation 4.4:
x

k
Cs
= 1 (1 k ) * e l
Cu

(4.4)

where k is distribution coefficient and if thermodynamic conditions are taken into


account it is k0 or if the experimental results are analysed than it is taken as keff; x is
distance of the crystal interface from the point at which first crystals froze, and l is the
length of the molten zone.
A central question when considering the use of fractional crystallization for refining
aluminium is the extent to which secondary constituents can be purged from the base
metal. The theoretical limits for some binary Al alloys are summarised in table 4.1.
In table 4.2 calculated distribution coefficients for cases of ternary systems are
presented based on the work of Boender et. al. [4]. The calculations were performed by
using programs ChemSage and ThermoCalc using data of the COST507 project [11].
Thermodynamic studies of the alloying systems studied in this chapter revealed that for
all but one alloying system, there is no difference between distribution coefficient of
binary and ternary systems [4]. Only for AlSi1.5Mn1 a significant decrease of silicon
removal efficiency is found.
All of the listed systems are eutectic. They are thus appropriate for refining by means of
fractional crystallization processes. The theoretical prospects are good for iron, silicon,
and copper, fair for magnesium, and moderate for manganese.

67

Chapter IV

Table 4.1. Equilibrium distribution coefficients for the binary systems of the alloys

studied in this chapter [4].


Alloy

Element System

Thermodynamic
distribution coefficient
k0 ()

AlSi6

Si

eutectic

0.13

AlFe1.6Si0.4 Si

eutectic

0.11

Fe

eutectic

0.03

AlCu4Mg1.5 Cu

eutectic

0.17

Mg

eutectic

0.45

eutectic

0.13

eutectic

0.62

AlSi1.5Mn1.1 Si
Mn

Table 4.2. Calculated distribution coefficients for some binary and ternary aluminium

alloying systems[4].
Alloy

Element System

calculated distribution
coefficient k0 ()

AlSi6

Si

eutectic

0.13

AlFe1.2Si0.3

Si

eutectic

0.1

Fe

eutectic

0.03

AlCu4.4Mg1.5 Cu

eutectic

0.17

Mg

eutectic

0.45

Si

eutectic

0.17

Mn

eutectic

0.62

AlSi1.5Mn1

Finally, it should be noted that the conditions during the actual refining operation will
be principally different from the quasi-static conditions that apply for the phase
diagrams and theoretical calculations.

4.3. Experimental Procedure

Model alloys AlSi6, AlFe1.6Si0.4, AlSi1.5Mn1.1, and AlCu4Mg1.5 were prepared.


Casting of the ingots was done in a crucible that was cooled from the outer side in order

68

Zone melting investigation of the aluminium scrap refining potential in fractional crystallization

to obtain the same solidification conditions along the ingot. Cylindrical specimens with
diameter 10 mm and length 250 mm were machined from these blocks. These test bars
were taken from the central region of the ingots with their axis in line with the casting
direction.
Zone melting is a technique by which a specific redistribution of alloying elements can
be attained in a bar sample by repeated melting and solidification. A schematic of
sample and molten zone is presented in the insert of figure 4.2. Here, a sample of the
test material is pulled in horizontal direction through a ring-shaped heater, the position
of which is fixed. By this local heating, a confined part of the sample will melt. By
moving the sample through the heater, this molten zone travels through the sample from
head to end. This procedure means that the material is re-solidified at a controlled
crystal growth rate equal to the pulling rate of the sample.
Pulling direction

Heater

Sample

Molten zone

Stopper

Inert
gas

Induction
coil
Tube
Crucible with
sample

Guide
Pulling
wire

Weight
Linear
motor

Figure 4.2. The zone-melting device.

In figure 4.2, a drawing is given of the device that has been used for the experiments.
The zone refiner consists of a quartz tube (600 mm long and 25 mm in diameter), a
water-cooled copper coil and a high-frequency inductor (power P12 kW, frequency
range 28 MHz), and an arrangement to move the tube through the coil (possible
pulling rates of 0.0420 mm/min). The alumina crucible, that contains the sample, is
placed in the tube. Design of the crucible is based on Pfann [9] and its geometry is such
that the samples are largely enclosed, preventing the molten zone from sagging. After

69

Chapter IV

sealing the tube with teflon stoppers, it is flushed with argon. A small flow of argon is
maintained during the experiment in order to prevent the formation of an aluminium
oxide layer around the aluminium melt. Figure 4.3 shows the zone melted sample. Only
the middle part of the cylindrical bars with 120 mm length was zone melted. The reason
for choosing relatively long ends is to keep the lateral heat transfer during the test more
or less constant and to minimise variations in the molten zone length, which at these
experiments was in the order of 20 mm. The AlSi6 alloy was pulled at a rate of 0.1-4
mm/min, while all other alloys were pulled at 0.1 mm/min pulling rate. The induced
heat, controlled by the power supply of the inductor, had to be adjusted to the alloy
composition such that full melting occurs in a restricted zone.

Figure 4.3. The zone-melting specimens, showing the position for chemical analyses,

and start and end of melting of the sample, dimensions are in mm.

After zone melting the chemical compositions at distinct locations within the sample
were measured and the three locations (head, middle, and end) are shown in figure 3.
These analyses were performed with optical emission spectroscopy (OES). The OESanalyses were done on cross-sections of the bars, since the variation in composition is
minimal perpendicular to the pulling direction. Concentrations were measured at the
head, the middle and the end location for each of the treated samples and are denoted
Ch, Cm and Ce respectively. Alloying contents in the untreated materials were also
measured and are designated as Cu.
The samples for microscopic analysis were cut and subsequently ground with a grinding
chapter, followed by polishing with a diamond powder (starting with 9 m and ending
up with 1 m powder size). After polishing, samples were etched by inserting in a

70

Zone melting investigation of the aluminium scrap refining potential in fractional crystallization

Keller-Wilcox etching reagent (10 cm3 HF 40 % + 20 cm3 concentrated HCl + 15 cm3


concentrated HNO3 in 955 cm3 distilled water) for 15 seconds. Micrographs were taken
with a Neophot 30 inverted stage optical microscope equipped with a Polaroid digital
camera model DMC1 with DC input of 6 V, 3 A, connected to a PC.
Line scanning electron microprobe measurements were performed employing
Wavelength Dispersive Spectrometry (WDS). The composition at each location of the
sample was determined after background correction. The obtained intensity ratios were
processed with the matrix correction program CITZAF [12]. The AlSi6 samples grown
at different growth rates were analysed using Electron Probe Micro-Analysis (EPMA)
in combination with a line scan in order to determine the content of Si in large grains.
The points of analysis were located along the line with increments of 4 m and involved
the elements Al and Si.

4.4. Results

4.4.1. Chemical Analyses


In tables 4.3 and 4.4, the results of the chemical analyses for a series of experiments are
summarised. These represent the average values for three tests. Only the elements
exceeding 0.01 wt % in concentration in the untreated materials were considered.
As is obvious from these results, a significant refining effect occurs for each of the
involved major alloying elements, except for highest growth rate for AlSi6 alloying
system. Though the concentrations for the (incidentally trapped) other elements are
much lower, purification is noticeable here too. In most cases, measured concentrations
are similar at the head and at the middle location; this means that the purified part
extends over at least half of the sample. The concentration at the end of the sample
represents the final concentration of liquid, which shows the degree of accumulation of
impurities at the end of zone melting of the samples.
Table 2 shows that purification efficiency for AlSi6 alloy decreases sharply with
increasing growth rate (G) and at a growth rate of 66.67 m/s it is negligible.

71

Chapter IV

Table 4.3. Chemical composition of AlSi6 alloy before and after zone melting for

various growth rates (G).


G (10-6m/s)

Element original

Head

Middle

End

Cu (wt %) Ch (wt %) Cm (wt %) Ce (wt %)


1.67

Si

6.14

2.85

2.44

10.1

6.67

Si

6.14

4.83

4.33

9.54

16.67

Si

6.14

5.28

5.04

8.09

66.67

Si

6.14

6.04

6.15

7.52

Table 4.4. Chemical composition before and after zone melting (G=1.6710-6 m/s).

Alloy

Element original

Head

Middle

End

Cu (wt %) Ch (wt %) Cm (wt %) Ce (wt %)


AlFe1.6Si0.4

Fe

1.59

0.86

1.22

2.30

Si

0.37

0.12

0.12

0.78

Sb

0.06

0.03

0.05

0.09

Cu

4.42

1.53

3.21

Mg

1.49

0.85

1.21

2.35

Fe

0.03

0.01

0.02

0.10

Si

0.03

0.01

0.02

0.07

AlSi1.5Mn1.1 Si

1.51

0.45

0.52

3.76

1.09

0.83

0.82

1.30

AlCu4Mg1.5

Mn

10.3

4.2. Microscopic examinations


Samples of the AlSi6 alloy, grown at different pulling rates, were further studied by
optical microscopy.
Figures 4.4 and 4.5 show an overview of the microstructures for three growth rates and
at the three locations within the sample. At the two lower growth rates, the morphology
at the head and in the middle is similar, whereas at the end the darker regions,
representing the silicon-enriched areas, are clearly more dominant. This confirms that
purification has occurred. At the highest growth rate, this effect is no longer visible,
although the chemical analyses show a minor purification. For the two lower growth

72

Zone melting investigation of the aluminium scrap refining potential in fractional crystallization

rates, the cell size decreases from about 0.2 to 0.1 mm and the amount of silicon
between the cells increases when going from head position to end position. The latter
effect can be better seen in figure 4.5. In particular for the higher growth rates, the
microstructures indicate that while crystals are growing as primary aluminium
dendrites, melt is entrapped at the boundaries. This silicon-enriched melt subsequently
solidifies as an interdendritic eutectic (where the silicon appears as flakes).
In order to investigate the effect of the growth direction on the obtained crystal
morphology, samples were also cut and prepared in longitudinal direction. In figure 4.6,
the microstructures at the head position of the bar are shown for two growth rates and at
two levels of magnification. At the lower growth rate, the crystals have grown as
columns in the pulling direction, as can be deduced from the long dark lines
representing the silicon-rich intercellular areas. At the higher growth rate, this effect has
almost disappeared, although by comparing figures 4.5 and 4.6 it can be seen that the
aluminium-rich cells are still elongated in the growth direction by about a factor two.
The zone melting samples of the other alloying systems, grown at 1.67 m/s, were also
studied. The micrographs of the three ternary alloying systems, taken in longitudinal
direction in a position where 40 % of the samples had been grown, are shown in figure
4.8. In the Al-Si0.4-Fe1.6 alloy it is found that only at the beginning of the zone melted
sample a flat interface appears and that later eutectic growth takes place and that a sharp
transfer from flat to eutectic growth was observed at the position where about 40 per
cent of the sample was crystallized. The other two systems grow with flat interfaces and
no change in their structure was observed with optical microscopy.

73

Chapter IV

position

end

mid

head

1.66

16.67

66.67

Figure 4.4. Micrographs of the Al-Si6 alloy at different growth rates and at different positions in the bar after zone melting. Line segment (see

insert) corresponds with distance of 0.1 mm.

74

Zone melting investigation of the aluminium scrap refining potential in fractional crystallization

position

en

mid

head

1.66

16.67

66.67

Figure 4.5. Same micrographs as in figure 5, but with higher magnification. Line segment (see insert) corresponds with distance of 25 m.

75

Chapter IV

1.66

16.67

Figure 4.6. Micrographs of the Al-Si6 alloy at different growth rates taken in

longitudinal direction (growth direction) after 30 % of the sample was crystallized


(maybe mention: after growth distance of xx mm). Line segments (see inserts)
correspond with distance of 0.1 mm (bottom) or 55m (top).

Al-Fe1.6%-Si0.4%

Al-Cu4%-Mg1.5%

Al-Si1.5%-Mn1.1%

Figure 4.7. Optical micrographs of above cited alloying systems, growth rate 1.67

m/s. Micrographs are taken along the longitudinal (growth) direction.


76

Zone melting investigation of the aluminium scrap refining potential in fractional crystallization

4.4.3. Electron Probe Micro-Analysis


The average values of the EPMA line scans in the inner part of the grains in the head of
the samples

(Ch,g),

are given in table 4.5. Table 4.5 clearly shows that the

concentration of Si in the grains increases with increasing growth rate. However, this
increase is by far smaller than the one found from chemical analyses presented in table
4.3.
Figure 4.4 shows the line scan for AlSi6 alloy grown at pulling rate 16.67 m/s. It is
seen that the concentration of Si in the inner side of cell is stable and its average value,
given in table 4.5, is 1.02 wt %.
Table 4.5. The average values of the Si content in AlSi6, determined from EPMA line

scans for various growth rates (G) Analysis is done in the middle of the grains at the
head position (Ch,g).
G (10-6m/s) 1.67 16.67 66.67
Ch,g (wt %) 0.82 1.02

1.45

AlSi6.14
1.400

Wt %

1.200
1.000
0.800
0.600
0.400
30

35

40

45

50

55

60

Position in m

Figure 4.8. The Si content in EPMA line scan of AlSi6.14 alloy grown at 16.67 m/s.

77

Chapter IV

4.5. Discussion

4.5.1. Comparison of the chemical analyses with equilibrium predictions


Figure 4.9 shows theoretical curves for single-pass zone melting of the AlSi6 alloying
system under thermodynamic equilibrium for various effective distribution coefficients.
The curves are calculated using Eq. (4.4) until the liquid zone reaches the eutectic
concentration. After reaching the eutectic concentration purification stops and the
solidified concentration becomes equal to the original concentration of the sample. At
the same time the microstructure will change to a dendritic structure containing -Al
and eutectic. The higher the effective distribution coefficient the longer it takes to reach
eutectic concentration in the molten zone, since more silicon is entrapped in the
solidified part.
7
6
5

CS in wt % Si

5
4
4
3
3
2

2
1

0
0

x/l

Figure 4.9. Silicon concentration after single-pass zone melting of AlSi6 alloying

system for: k0 = 0.13 (curve 1) and keff values of 0.2, 0.3, 0.4 and 0.5 (curves 2, 3, 4,
and 5, respectively).

Based on equation 4.4 the thermodynamic and experimental results of alloys studied in
this chapter are shown in figures 10-13. Pfann principle [9], which assumes no diffusion
in solid and complete diffusion in liquid, was used for calculating theoretical curves. It

78

Zone melting investigation of the aluminium scrap refining potential in fractional crystallization

is clearly seen in figure 4.10 that theory predicts that for AlSi6 only 1.5 molten zones
pass and than steady state is achieved and no purification takes place in the rest of the
sample. This is due to the fact that liquid reaches its eutectic composition and the
concentration of the metals formed is equal with the concentration of the starting
material that melts on the other side of the zone. The experimental purification of AlSi6
grown at growth rate (G) 1.67 m/s shows that it disagrees with theoretical predictions.

AlSi6
7
6

CS in wt %

5
4
3
2
1
0
0

x/l

Figure 4.10. Curves for single-pass zone melting, showing theoretical (lines) and

experimental () solute concentration in crystals versus distance in zone lengths (x/l)


from beginning of charge, for AlSi6 alloying systems.

The experimental results show a decrease on concentration going from beginning to


middle of the sample (after 3 molten zone lengths were passed). An explanation for this
phenomenon could be that in the beginning crystal interface is not stable due to
nucleation of first crystals and therefore more impurities are entrapped. The results of
the line scans of the cells at the head of the samples show close agreement with
predictions, with lower than expected concentration for the first two growth rates and
slightly higher for the highest growth rate.
Another alloy that reaches its steady state even faster is AlFe1.6, only the first half of
the molten zone will get purified (figure 4.11). The rest of the alloying elements get
purified for all the 3 molten zones that were passed (figures 4.12-4.13). Except for

79

Chapter IV

AlCu4.4 alloy, which behaves as expected form theory, but with lower purification
efficiency, all the other alloys show non or very small concentration difference on
crystals along the sample. It is clearly shown that purification efficiency for most of the
alloying elements was significantly lower than theoretically predicted and therefore no
eutectic concentration was reached except for AlFe1.6 alloy. The best results compared
to theoretical predictions are achieved for Mg (figure 4.12).

AlSi0.4Fe1.6

CS in wt %

2
1.5
1
0.5
0
0

x/l

Figure 4.11. Curves for single-pass zone melting, showing theoretical (full line for Si

and dashed line for Fe) and experimental ( for Fe and for Si) solute concentration
in crystals versus distance in zone lengths (x/l) from beginning of charge, for
AlSi0.4Fe1.6 alloying systems.

CS in wt %

AlCu4.4Mg1.5
3.5
3
2.5
2
1.5
1
0.5
0
0

x/l

Figure 4.12. Curves for single-pass zone melting, showing theoretical (full line for Cu

and dashed line for Mg) and experimental ( for Mg and for Cu) solute

80

Zone melting investigation of the aluminium scrap refining potential in fractional crystallization

concentration in crystals versus distance in zone lengths (x/l) from beginning of charge,
for AlCu4.4Mg1.5 alloying systems.
Roughly 24 to 70% of the major alloying elements is removed from the material at the
head of the sample. However, the thermodynamic limits are, in most cases, and in
particular for iron, not obtained by far. A reason for this is that the tests were carried out
at relatively high growth rate. A further explanation may be that of interference between
alloying elements in case of the ternary alloys which may have a negative effect on
purification efficiency.

CS in wt %

AlSi1.5Mn1.1
1.2
1
0.8
0.6
0.4
0.2
0
0

x/l

Figure 4.13. Curves for single-pass zone melting, showing theoretical (full line for Si

and dashed line for Mn) and experimental ( for Mn and for Si) solute concentration
in crystals versus distance in zone lengths (x/l) from beginning of charge, for
AlSi1.5Mn1.1 alloying systems.

In conclusion one can say that experimental conditions applied to these alloying systems
were fare from equilibrium and therefore they disagree with theoretical predictions.

4.5.2. Comparison of results for AlSi6 alloy with non equilibrium BPS-model
The AlSi6 alloy was processed in the zone refiner at different pulling rates (i.e., growth
rates) in order to screen for the optimal operating conditions and to establish the validity
of the BPS-model. The results are plotted according to the BPS-model (equation 4.2) in
figure 4.14. As an approximation, the densities of solid and liquid aluminium are taken
equal. Clearly, the model, which predicts a straight line with a slope corresponding to

81

Chapter IV

the mass transfer coefficient kd and an intercept value k=k0 for G=0, is not correct when
taking into account the overall concentration of crystals. The slope of the curve changes
from an apparent mass transfer coefficient kd,1=810-7 m/s at low growth rates to
kd,2=310-5 m/s at higher values (say for G>510-6 m/s). Compared to this, a typical value
of 7.710-6 m/s was found for zone melting of phenol [13]. BPS model fits very well
when concentration of Si in grains is taken into account. It shows a strait line which
intercepts the y axes at value k=k0 for G=0.

ln[1/keff-1]

2
1
0
-1
-2
-3
-4
0

10

20

30

40

50

60

70

Growth rate G, 10 m/s

Figure 4.14. BPS-representation for zone-melting experiments at different pulling rates

(alloy AlSi6), for average concentration in crystals () and for concentration in the
grains (). Stipulation with dashed line shows where according to prediction of BPSmodel, the line should touch the y axes at k0.
Figure 4.14 clearly indicates that the BPS-model cannot be used to accurately describe
the relation between the crystal growth rate and the distribution coefficient. Chang at al.
[14], who analyzed the crystallization of the benzene-hexadecane and used equation 4.2,
found that there were two regions of different linearity over the range of the growth
rates measured. After analysis of the interface morphology, they found out that each
region is associated with a distinguished crystal interface. The first region of linearity
was at a growth rate up to 1.2 cm per hour and corresponded to a cellular structure,
while for higher growth rates the interface was dendritic. As can be clearly seen from
the micrographic examination of the samples analyzed in this study, similar findings
were the case hear as well. Trapping of the solute in the intercellular or interdentritic
regions at the interface appears to be the main reason for such anomalous behavior of
82

Zone melting investigation of the aluminium scrap refining potential in fractional crystallization

this alloying system. BPS treated diluted solutions, assuming flat interface, which
clearly was not the case with our alloying system. Stepanski and Schfer [15] showed
that the degree of purification in layer growth process is strongly dependent on the
concentration of impurities in the system. They show the effect of the initial impurity
concentration in purification efficiency for the system Phenothiazine (PTZ) in acrylic
acid. It is clear from their obtained results that separation is significantly less effective
at higher initial impurity concentrations. This can be manly attributed to the quantitative
increase of the impurity inclusions due to the increased impurities, resulting in a crystal
layer with more entrapped and adhered liquid.
Since a proper model is essential for the design of a layer crystallization process, these
findings suggest that it is better to develop an empirical model based on the obtained
experimental results rather than relying on the theoretical models.

4.6. Conclusions

Zone melting is a relatively simple and useful experimental technique to assess the
ability to refine distinct alloying constituents from complicated alloy systems (like
industrial kinds of scrap aluminium). Regarding the obtained purification in the
experiments, results basically comply with the estimates based on the binary alloy phase
diagrams, although there are known distinct deviations from other experimentally
achieved results. Chemical analyses reveal that purity of the samples on the head and on
the middle positions are almost the same, which reveals that purification takes place in
more than the half of the original samples.
Microscopic examination of the AlSi6 alloy reveals that while aluminium cells grow,
the impurities that rich eutectic concentration are trapped between them. The
AlSi0.4Fe1.6 alloying system micrographs show a flat interface till about 40 per cent of
the sample is crystallized and later a eutectic growth resumes. The other two systems
studied in this chapter grow with flat interface and no change in their morphology was
observed by microscopy.
The BPS-model does not satisfactorily correspond with the experiments, and needs
extension to obtain a predictive tool for designing refining processes based on layerbased fractional crystallization. The microstructures of zone-refined samples indicate

83

Chapter IV

that an improved purification model should include the amount of eutectic trapped
between the cells.

4.7. Acknowledgements

Reported investigations were carried out as a part of the research project Sustainable
Technology for the Reclamation of High-grade Aluminium from Scrap, which is
undertaken in the context of the Dutch R&D stimulation programme on Economy,
Ecology and Technology (contract number EETK98026). Concerning the experiments,
the assistance of Mr C.M.S. van der Zwet Slotenmaker and Mr P.G. Commandeur at
Corus Group plc RD&T (casting of model alloys) and Mr C.C.J. Kaasschieter and Mr
M.H.F.M van Hout at TNO Institute of Industrial Technology (zone-melting tests) are
acknowledged.

84

Zone melting investigation of the aluminium scrap refining potential in fractional crystallization

4.8. References

1.

W.H. Sillekens, D. Verdoes and J.A.F.M. Schade van Westrum, Refining


Aluminium Scrap by means of Fractional Crystallization: Technical
Feasibility, 4th ASM International Conference and Exhibition on the Recycling
of Metals, ASM Europe, 1999, p. 105.

2.

T. Sotome and M. Ohtaki, 1998, "Application of Fractional Crystallization for


Refining of Molten Aluminium Scrap", Conference Proceedings of ICAA 6, vol.
1, Ed. T. Sato et al. Tokyo, Japan: The Japan Institute of Light Metals, 351.

3.

A.I. Kahveci and A. Unal, 2000, "Refining of a 5XXX Series Aluminum Alloy
Scrap by Alcoa Fractional Crystallization Process", Proceedings of the Fourth
International Symposium on Recycling of Metals and Engineered Materials, Ed.
D.L. Stewart Jr., J.C. Daley, and R.L. Stephens, Warrendale PA, USA: TMS, 979.

4.

W. Boender, C.J. Waringa, G.P. Krielaart, A. Folkertsma, and D. Verdoes, 2002,


Refining Aluminium Scrap With Fractional Crystallization: Assessing Its
Feasibility With Thermodynamics, Proceedings of the TMS Annual Meeting &
Exhibition, Light Metals 2002. Ed.Schneider W., Seattle, Washington, USA,
TMS,1061.

5.

G.J. Arkenbout, 1995, Melt Crystallization Technology. Lancaster PA USA:


Technomic Publishing Company Inc.

6.

P.J. Jansens & M. Matsuoka, 2000, Melt Crystallization, in: Encyclopedia of


Separation Science, Eds. Wilson I.D., Adlard E.R., Cook M., and Poole C.F.,
Academic Press, San Diego, New York London, Sidney, Tokyo, 966.

7.

J.D. Esdaille and G.W. Walters, 1975, "Lead Purification by Crystallization and
Reflux", Proceedings of the Aus. I.M.M. conference, South Australia, 187.

8.

W.G. Pfann, Trans. AIME, 194 (1952) 747.

9.

W.G. Pfann, Zone Melting, John Wiley & Sons Inc., (2nd edition), New York, NY,
U.S.A., 1966.

10. J.A. Burton, R.C. Prim and W.P. Slichter, The Journal of Chemical Physics, 21
(1953) 1987.
11. P. Spencer, definition of thermochemical and thermophysical properties to
provide a database for the development of new light alloys, Proceedings of the

85

Chapter IV

final workshop on COST 507, Volumes 1, 2 and 3, Vaals, The Netherlands 9-11
March 1997, EUR 18499, European communities, 1998.
12. J.T. Armstrong, Quantitative elemental analysis of individual microparticles with
electron beam instruments. In Electron Probe Quantitation, Heinrich, K.F.J.,
and Newbury, D.E. (eds.), Plenum Press, (1991) 261-315.
13. P.J. Diepen, Cooling Crystallization of Organic Compounds: Processes, Purity
and Permeability, Ph.D. Thesis, Delft University of Technology, 1998.
14. C. Chang, D.Irving, and B. Kyle, AICHE-Journal, 13 (1967) 739.
15. M. Stepanski and E. Schfer, In Course on:Theory and Application of melt

Crystallization, 28th 30th September 2000, Martin Luther University HalleWitwnbwrg, Halle (Saale), Germany.

86

Zone melting investigation of the aluminium scrap refining potential in fractional crystallization

87

CHAPTER V
5. ALUMINIUM SCRAP REFINING WITH SOLID LAYER
FRACTIONAL CRYSTALLIZATION1

Abstract
Primary aluminium production is an energy intensive process. With an increase of the
aluminium use in car manufacturing a recycling technological process is imperative.
The energy use for recycling is only a fraction of the energy needed to produce primary
aluminium. In this chapter a fractional crystallization process for aluminium recycling
based on static layer growth will be described. Static layer crystallization experiments
showed that for AlSi7.2 alloying systems, typical growth rates of 1 to 2 m/s are
optimal for aluminium scrap purification. In order to achieve the preset purification
criteria for the pure product, one or more re-crystallization steps are required.
Depending on the impurity concentration, the crystallization temperature varies from
600 to 660 C. Microscopic examination of the crystal samples reveals that, while
crystals are growing as primary aluminium dendrites, impurity enriched melt is
entrapped at the boundaries. Line scans reveal that in the crystal cell area purification is
in line with thermodynamic predictions, while intercellular area is enriched with
impurities and intermetalics, which have an adverse effect in purification efficiency of
selected aluminium alloys.

Part of this work was published by Mehmetaj B., Bruinsma O.S.L., Kool W.H., Jansens P.J., and
Katgerman L.: Proceedings of the 15th International Symposium on Industrial Crystallization (ISIC-15),
Ed. Chianese, A., Sorrento, Italy, Chemical Engineering Transactions, Volume 1, 2002, p.879 884.

Chapter V

5.1. Introduction
The manufacturing industries are focusing more and more on processes that address an
increased environmental awareness on the part of the public and stricter regulations for
chemical disposal and pollutant emission by the governments. As a result new
environmental friendly and energy efficient technologies are emerging.

Such an

example is melt crystallization technology, which provides a unique separation potential


for products that require high purity applications in addition to environmental and
energy efficiency benefits. One such technology development is recycling aluminium
scrap by means of fractional crystallization.
Production of aluminium from bauxite is an energy intensive process. For production of
1 kg aluminium is required as much as 14 kWh of electrical energy, whereas for remelting of aluminium scrap a fraction of that energy is needed. The energy required for
recycling of aluminium scrap is only 5-10 percent of the energy needed for producing
primary aluminium. Given these facts, aluminium recycling can lead to significant cost
savings.
A common practice in aluminium recycling is the re-melting of aluminium scrap and
reuse for the foundry alloys. In cases, where the alloying contents are higher than
requirements for producing a specific alloy, the molten scrap is diluted with primary
aluminium till the required chemical composition is achieved. In the short term there is
no threat to this practice. However, with a continuous increase of the aluminium
consumption the amount of aluminium scrap that needs to be recycled cannot be
processed in this way. Therefore a closed loop recycling is an imperative. Since
conventional metal refining cannot apply to aluminium, fractional crystallization is seen
as the best alternative.
Melt crystallization is a separation process by which the fractional separation is effected
by directional crystallization from the melt. Fractional crystallization is applied for both
ultra purification of organic compounds [1-2], and metals [3]. Historically, the first
application of fractional crystallization for metals was the invention of the Pattison
process for extraction of silver from lead in 1829 [4]. It was followed by the
development of the zone melting technique in the 1950s by Pfann [5].
Besides for lead purification, fractional crystallization is also used for production of
super pure aluminium from virgin metal (see reference [3] for an overview). In this
project an attempt is made to develop a fractional crystallization process for aluminium
88

Aluminium scrap refining with solid layer fractional crystallization

scrap recycling. The zone melting results of the alloys studied in this project were
published by Sillekens et al. [6], while a thermodynamic study of the alloys investigated
in this project is provided by Boender et al. [7]. Similar projects were carried out in
Japan [8] and in the USA [9].
In this chapter a layer based fractional crystallization process is described. A
background on application of fractional crystallization in metallurgy is given in section
5.2. In section 5.3 the experimental procedure and equipment description is provided.
The chemical analyses, EPMA line scans, and micrographic description of the studied
alloys grown at different growth rates are provided in section 5.4.

5.2. Theory
In evaluating the performance of a fractional crystallization process, an indicator called
the effective distribution coefficient, keff, is used. Relating to a mixture of the
components A (here: aluminium) and B (any other element), it is defined as:

k eff =

CS
.
CL

(5.1)

CS and CL are the concentrations of component B in the solidified fraction and in the
liquid material respectively. The effective distribution coefficient depends on the
thermodynamic properties of the system and on the process kinetics. The distribution
coefficient in the equilibrium condition can be derived from the phase diagram of the
concerned system and is called the thermodynamic distribution coefficient, k0.
A central question when considering the use of fractional crystallization for refining
aluminium is the extent to which secondary constituents can be purged from the base
metal. The theoretical limits for some binary Al alloys are summarised in table 5.1.
In table 5.2 calculated distribution coefficients and T values in case of studied binary
and ternary systems are presented based on the work of Boender et al. [7]. The
calculations were performed by using programs ChemSage and ThermoCalc using data
of the COST507 project [10]. Thermodynamic studies of the alloying systems studied in
this chapter revealed that for all but one alloying system, there is no difference between
the distribution coefficients of binary and ternary systems [7]. Only for AlSi1.5Mn1 a
significant decrease of silicon removal efficiency is found.
89

Chapter V

Table 5.1. Thermodynamic limits on the purification of secondary constituents from


aluminium by means of fractional crystallization (single-step operation).
Element

System

Si

eutectic

Fe

eutectic

Cu

eutectic

Mn

eutectic

Mg

eutectic

Maximum temperature

Thermodynamic

range T (C)

distribution coefficient k0 ()

83

0.13

5.5

0.03

112

0.17

2.5

0.62

210

0.45

Table 5.2. Calculated distribution coefficients for some ternary alloys and temperature
ranges of the semi-solid region for the critical alloying component. [7].
Alloy

Element System Temperature


range T (C)

calculated
distribution
coefficient k0 ()

AlSi6

Si

eutectic 40

0.13

AlFe1.2Si0.3

Si

eutectic

0.1

Fe

eutectic 3

0.03

eutectic 99

0.17

Mg

eutectic

0.45

Si

eutectic

0.17

Mn

eutectic 2

0.62

AlCu4.4Mg1.5 Cu

AlSi1.5Mn1

With reference to temperature control for such processes, the maximum temperature
range T is included in the table 5.2. It is taken as the temperature interval of the semisolid state in which the purification of the alloying system occurs (i.e., TlTE), where Tl
is liquid temperature of the given alloying system (in table 5.1 it is liquid temperature of
pure aluminium), and TE is eutectic temperature of a given binary system. The limiting

90

Aluminium scrap refining with solid layer fractional crystallization

factor in ternary alloys is reaching the eutectic composition (table 5.2), after which no
further purification will take place during solidification.
All of the listed systems are eutectic. They are appropriate for refining by means of
fractional crystallization. The theoretical prospects are good for iron, silicon, and
copper, fair for magnesium, and moderate for manganese. Besides this, process control
will be critical for some cases: the available temperature interval is very small for iron
and manganese.

5.3. Experimental
The static layer crystallizer consists of a Nabertherm K4/10 tilting furnace and a
Eurotherm 2604 temperature controller. The furnace has a 4-liter crucible of ISOgraphite, with the maximum chamber temperature of 1273 K, designed for melting
aluminium. Thermocouples are made of NiCr-Ni, type K. A schematic drawing of the
static layer crystallizer is depicted in figure 5.1 and the equipment for this process is
given in figure 5.2.
The temperature controller is designed in such a way that during the melting stage the
temperature is controlled using the thermocouples attached on the outside wall of the
crucible, and during the crystallization stage the temperature is controlled with the
thermocouple inserted in the aluminium melt. Both stages are PID controlled, with
automatic switch from melting to crystallization PID mode. Eurotherm 2604
incorporates a self-correcting input circuit to preserve the instrument calibration
accuracy. Temperature configuration is achieved either via the front panel interface or
by using the Eurotherm iTools configuration package, which runs under the Windows
2000 operating systems.
The alloy of aluminium with 7.2 percent silicon (AlSi7 corresponding to its nominal
composition according to the EN AW system) is prepared by addition of a certain
amount of ALSi50 to the technically pure primary aluminium melt. The ternary
aluminium alloys described in this chapter were prepared by adding of a pre-calculated
amount of master alloys to the commercially pure aluminium melt. The cold finger is
inserted to the prepared alloy and the temperature is lowered to the melting temperature
of the alloy by vigorous air blowing to the cold finger. After the melt temperature is left
to stabilize for some time the cooling rate is adjusted by airflow to achieve the desired

91

Chapter V

growth rate. The growth rate is determined by dividing the crystal layer thickness by the
time it took to grow it.

Argon

5 cm

TC

10 cm

Crystal layer

Cooling
air

20 cm

Figure 5.1. Static layer crystallizer for aluminium scrap refining.

Cold
finge
Crystallization
oven
Crystal
Temperature layer
controller

Figure 5.2. Equipment of static layer crystallizer for aluminium scrap refining.
The samples for chemical analyses were cut to include all the layers of the crystal. The
AlSi7.2, AlFe1.3Si0.3, AlCu4Mg1.4, and AlSi1.5Mn1 samples were dissolved in a
mixture of 10 ml water, 10 ml HCl, and 6 ml HF. A closed Teflon vessel was used to

92

Aluminium scrap refining with solid layer fractional crystallization

dissolve the samples in a microwave oven. Chemical analyses of the above mentioned
alloys, except for Mn, were carried out using atomic absorption spectroscopic technique
(Perkin Elmer 2380(A)). The Mn analysis of the alloying system AlSi1.5Mn1 were
performed by Inductively Coupled Plasma Optical Emission Spectrometry (ICP-OES)
technique (Perkin Elmer plasma 40). The overall error of chemical analyses is 3 %.
The samples for microscopic analyses were cut and subsequently grinded with grinding
chapter fallowed by polishing with diamond powder (starting with 9 m and ending up
with 1 m powder size). The micrographs were taken in a Neophot 30 inverted stage
optical microscope equipped with a Polaroid digital camera model DMC1 with DC
input of 6 V, 3 A, connected to a PC.
The Electron Probe Micro Analaysis (EPMA) were performed with a JEOL JXA 8900R
microprobe using an electron beam with energy of 15 keV and beam current of 20 nA
employing Wavelength Dispersive Spectrometry (WDS). The composition at each
analysis location of the sample was determined using the X-ray intensities of the
constituent elements after background correction relative to the corresponding
intensities of reference materials. The points of analysis were located along a line with
increments of 4 m and involved the elements Al, Cu, Mg, Fe, Mn, and Si. The thus
obtained intensity ratios were processed with a matrix correction program CITZAF
[11]. The counting errors and detection limits are given in table 5.3 for each spectral
line used in the quantitative analysis.
Table 5.3. The counting error (one sigma) and detection limits for the measured X-ray
lines.
Element X-ray Line

Counting error Detection Limit

(Energy keV) (%)

(p.p.m.)

Al

K (1.487)

0.05

10

Cu

K (8.046)

0.05

51

Mg

K (1.254)

0.02

11

Mn

K (5.898)

0.05

37

Si

K (1.740)

0.02

16

Fe

K (6.403)

0.06

40

93

Chapter V

Scanning electron microscopy in backscattered electron configuration was used to make


pictures of the locations of areas in which the EPMA analyses were performed. The
same conditions as with the Quantitative Analysis were used.

5.4. Results and discussion


In order to establish the potential of refining aluminium scrap by means of fractional
crystallization processes, AlSi7.2, AlFe1.3Si0.3, AlCu4Mg1.4, and AlSi1.5Mn1 were
selected. The first alloy was included as it represents a rather basic and comprehensible
model system. The alloy was crystallized at different growth rates in order to find the
optimum growth rate.

Other alloying systems were chosen as they represent

concentrations of the aluminium scrap streams in the Netherlands. The obtained crystal
layers were analysed with chemical analyses, EPMA line scans, and their micrographic
examination was carried out.
5.4.1 Chemical analyses
The results of the chemical analyses of the layer growth experiments for AlSi7.2 alloy,
carried out at different growth rates (G), presented as overall distribution coefficient
(koa= Ccr/C0), which is defined as a concentration of the impurities in the crystal layer
(Ccr) versus impurity concentration of the starting alloy (C0), and average distribution
coefficient (kcr, equation 5.2), are shown in figure 5.3. The koa expression gives the
overall partition of impurities and does not take into account the changing concentration
of impurities in the liquid. In order to account for that, we consider an initial mass of
liquid m0, with impurity concentration C0, being partially crystallized to give a reduced
mass of liquid mr, with an impurity concentration Cr. Since effective distribution
coefficient (keff) can be considered as a constant, the mass and component balance for a
differential balance of crystallized mass can be integrated to give an average distribution
coefficient kcr [1]:

k cr = 1 +

log(C r / C 0 )
log(mr / m0 )

(5.2)

94

Aluminium scrap refining with solid layer fractional crystallization

AlSi7.2
0.65
0.6

0.55
0.5
0.45
0.4
0.9

1.1

1.3

1.5

1.7

1.9

2.1

G in m/s

Figure 5.3. Overall distribution coefficients (koa, ) and average distribution coefficient
(kcr, ) versus growth rate (G) for AlSi7.2 alloy.
In practice the main advantage of the model described by equation 5.2 is that it gives an
expression of the purification achievable in a single stage depending on the amount of
the original liquid remaining un-crystallized. It also gives a more accurate value to
compare with theoretical predictions as given by thermodynamic partition coefficient
(k0).
Considering that about 40 % of the original material was crystallized the overall
impurity removal efficiency ( = 1 koa) of the layer growth process for AlSi7.2
alloying system is significant and varies from 37 % for the G = 2 m/s to50 % For G=1
m/s. In order to compare with theoretical possibilities we take the average distribution
coefficient for G = 1 m/s, which shows a removal efficiency of 57 % and compare it
with the thermodynamic possibilities. The impurity removal efficiency for this growth
rate is significantly smaller than thermodynamically possible (87 %). The reason for this
is that the tests were carried out at finite growth rates. Chemical analyses of the alloy
grown at 16 m/s revealed that no purification happened.
The impurity removal efficiency for the alloy grown at 1.7 m/s by zone melting
technique [6] is between the one observed for the same alloy grown at 1 m/s and 1.7
m/s by layer growth process. However, comparing the purification efficiency of the
layer growth process and zone melting technique for the 1.7 m/s growth rates, the
purification efficiency of zone melting is about 5 % better. There are two main possible

95

Chapter V

explanations for that: (i) incorporated impurities due to mother liquor adhering at the
end of the crystallization process were not removed effectively (no additional washing
was applied); and (ii) Zone melting experiments were carried out by using induction
heating, which besides heating provides a significant mixing effect.
Results of the chemical analyses of the layer growth experiments of the AlFe1.3Si0.3,
AlCu4Mg1.4, and AlSi1.5Mn1, presented in weight percent, overall partition
coefficient, and average partition coefficient, are shown in table 5.4.
Table 5.4. Chemical composition of the samples before and after layer growth.
Alloy
AlSi0.3Fe1.2
AlCu4Mg1.5
AlSi1.5Mn1

Element

Untreated

Crystals

C0/w%

Ccr/w%

koa

kcr

Si

0.33

0.18

0.54

0.31

Fe

1.33

0.71

0.53

0.23

Cu

4.02

2.68

0.66

0.35

Mg

1.39

0.74

0.53

0.33

Si

1.46

0.75

0.51

0.48

Mn

1.02

0.86

0.84

0.7

The applied growth rate (G) for all the alloying systems presented in table 5.2 was G=2
m/s. The impurity removal efficiency is significantly smaller than thermodynamically
possible. The overall purification efficiency of the iron and silicon on AlFe1.3Si0.3 are
almost the same, which based on the binary phase diagram is not expected. The
purification efficiency of iron was negatively effected by presence of silicon, while the
purification efficiency of silicon was positively effected by presence of iron if a
purification efficiency of the binary alloy is taken into account. However, from the
average distribution coefficient it is shown that partition of iron is higher than that of
silicon. In this case 40 % of the original material was crystallized and concentration of
iron in remaining liquid was almost 2 wt%, which is higher than eutectic concentration
of AlFe binary alloy (1.9 wt %). This shows that no more than 40 % of the original
material of this system can be crystallized in layer crystallization process at this growth
rate.
A high purification efficiency of magnesium on AlCu4Mg1.4 alloying system was
observed. This can be attributed to two possible effects: (i) evaporation of the

96

Aluminium scrap refining with solid layer fractional crystallization

magnesium; or (ii) loss of the magnesium to dross formation. These possible


explanations can be validated by checking the net loss of magnesium when doing mass
balance of original material with its concentration in crystals and remaining melt. Also
for this alloy 40 % of the original material was crystallized.
The purification of Mn and Si in AlSi1.5Mn1 alloy is as expected, with Mn purification
close to thermodynamic possibilities. As predicted from thermodynamic calculations for
these alloy [7] the removal efficiency of silicon is negatively affected. For this alloy a
quarter of the original material was crystallized.
In general, the impurity removal efficiency for the alloying systems described above is
in line with the results obtained by zone melting technique [6]. The impurity removal
efficiency is negatively affected by increasing concentration of alloying elements in the
starting material. This phenomenon can be clearly seen by looking on example of
silicon on different alloying systems described in this chapter (see table 5.2 and figure
5.2). Similar findings were found for the system Phenothiazine (PTZ) in acrylic acid
[12]. The chemical analyses for alloying systems presented in this work reveal that
significant purification is achieved. However, if an effective removal of the silicon from
aluminium scrap is to be achieved a repetition of the crystallization process has to be
performed.
5.4.2. Microscopic Examinations
In order to investigate the effect of the growth direction on the obtained microstructure,
samples AlSi7.2 alloy were cut and prepared in longitudinal direction. In figure 5.4a the
crystal grown at 1 m/s, cut in the growth direction, horizontal view is presented. The
left side is close to the cold finger and growth occurs towards the right side. It is seen
that crystals have grown as columns in the growth direction, as can be deduced from the
long dark lines representing the silicon-rich areas. Similar effects were observed for the
samples of the same alloy grown at 1.7 m/s by the zone melting technique [6].
Figure 5.4 shows an overview of the microstructures for 4 growth rates. The
microstructures suggest that while crystals are growing as primary aluminium cells,
melt is entrapped at the boundaries. This silicon-enriched melt subsequently solidifies
as an inter-cell eutectic (where the silicon appears as flakes). This observation is in line
with results of chemical analyses which show that purification has occurred except for

97

Chapter V

the highest growth rate. At the highest growth rate, this effect is no longer visible. The
cell size decreases with increasing growth rate. For the two middle growth rates, the cell
size decreases in going from about 0.4 to 0.3 mm and the amount of silicon between the
cells increases. Looking in comparison with the zone melting results the cells are twice
as large, for the same growth rate. However the amount of the inclusions between the
cells is much higher for the layer grown samples.

a)

b)

c)

d)

Figure 5.4. Optical micrographs for AlSi7 alloy from layer growth experiments: a)
grown at 1m/s seen in the direction of growth; b), c) and d) cross-sections of the
samples grown at respectively 1.6, 2, and 16 m/s.
The microstructure of the first three micrographs reveals that impurities are mainly
concentrated between the cells. Due to the low temperature gradients for the aluminium
layer, these impurities are entrapped as a liquid metal between the cells at the time when
crystallization stage is finished and may be removed by sweating or washing postpurification processes. If all the intercellular material would be removed by sweating,

98

Aluminium scrap refining with solid layer fractional crystallization

the purification efficiency would be increased considerably by loss of some 10 % of


crystal mass, i.e. intercellular material, for G = 2 m/s.

AlFe1.3Si0.3

AlCu4Mg1.4

AlSi1.5Mn1
Figure 5.5. Optical Micrographs for layer grown samples of the above cited alloying
systems, growth rate 2 m/s (longitudinal sections on the left; cross-sections on the
right). The line in the picture denotes 0.1 mm.

99

Chapter V

In figure 5.5 are presented micrographs of AlFe1.3Si0.3, AlCu4Mg1.4, and AlSi1.5Mn1


crystals grown at growth rate G= 2 m/s, cut in the growth direction (left side) and
cross-sections (right side). It is seen that crystals have grown as columns in the growth
direction. The microstructures suggest that while crystals are growing as primary
aluminium cells, impurity rich melt is entrapped at the boundaries. The cell size for all
the alloying systems presented in figure 5.1 is much larger than for AlSi7.2 alloy. The
cell size of AlSi1.5Mn1 is particularly large, which can be due to the lower initial
impurity concentration of this alloy compared with AlSi7.2. The other two alloying
systems show similar features. The microstructure of the first three micrographs reveals
that impurities are mainly concentrated between the cells. Due to the low temperature
gradients for the aluminium layer, these impurities are entrapped as a liquid metal
between the cells at the time when crystallization stage is finished and may be removed
by sweating or washing processes.

5.4.3. Line scans


The line scans for alloys studied in this project are presented in figures 5.5 - 5.8. The
Applied growth rate (G) for all the alloying systems was G=2 m/s. The impurity
removal efficiency on the cells is in line with thermodynamic predictions, except for
Mg whose concentration in cells is even lower than that.
The line scan for AlSi7.2 alloy is presented in figure 5.6. It is clearly shown that
concentration in the middle of the cell is close to the thermodynamically predicted one
for crystals. The concentration of Si close to the intercellular region increases and
within crystal it reaches maximum eutectic solubility of Si in Aluminium, which is
shown by spills in the curve. The intercellular region has eutectic concentration.

100

Aluminium scrap refining with solid layer fractional crystallization

AlSi7.2
4

Si (wt-%)

3
2
1
0
0

100

200

300

400

500

600

700

position in m

Figure 5.6. The Si content in EPMA line scans of AlSi7.2 alloy cells grown at growth
rate 2 m/s. The measurement was taken at a marked line in figure above the line scan.
The line scans for both Si and Fe of AlFe1.3Si0.3 alloying system presented in figure
5.7 are within the thermodynamic predictions. The significant oscillations of the
concentration lines, particularly for Fe, are due to the counting error for these elements
(see table 5.3). The concentration of Fe in crystals is lower than the counting error and
therefore shows an indication of concentration rather than an accurate concentration
line. In contrast to line scans, the overall purification efficiency of the iron and silicon
on AlFe1.3Si0.3, as shown in table 5.4, are almost the same. The Purification efficiency
of iron is negatively affected by intermetalic compound formations that were found out
to be present on the intercellular regions and will be discussed in the next section.

101

Chapter V

AlFe1.3Si0.3

wt %

Iron

Silicium

0.30
0.25
0.20
0.15
0.10
0.05
0.00
0

100

200

300

400

500

600

Position in m

700

800

900

1000

Figure 5.7. The Si and Fe content in EPMA line scans of AlFe1.3Si0.3 alloy cells grown
at growth rate 2 m/s. The measurement was taken at a marked line in figure above the
line scan.
In figure 5.8 the line scans for Mg and Cu of the AlCu4Mg1.4 alloying system are
presented. The purification efficiency of Cu in crystals is as expected from
thermodynamic predictions. It is notable the high purification efficiency of Mg on
AlCu4Mg1.4 alloying system. Similar findings for this alloy were reported for overall
purification efficiency presented in section 5.4.1 and the same reasoning for increased
purification efficiency could be applied.

102

Aluminium scrap refining with solid layer fractional crystallization

AlCu4Mg1.4

wt %

Copper

Magnesium

2.0
1.8
1.6
1.4
1.2
1.0
0.8
0.6
0.4
0.2
0.0
500

400

300
200
Position in m

100

Figure 5.8. The Cu and Mg content in EPMA line scans of AlCu4Mg1.4 alloy cells
grown at growth rate 2 m/s. The measurement was taken at a marked line in figure
above the line scan.
A remarkable result is that Mn concentration on cells is almost the same like overall
concentration on the crystals. This is due to the higher content of crystal cells in
comparison with intercellular material and lower concentration of Mn in the
intercellular area compared to other alloying elements studied in this work. A peak on
Si content is present in figure 5.9, which could be due to the invisible grain boundary
presence.

103

Chapter V

AlSi1.5Mn1

wt %

Manganese

Silicium

1.00
0.90
0.80
0.70
0.60
0.50
0.40
0.30
0.20
0.10
0.00
0

50

100

150

200

250

Position in m

Figure 5.9. The Si and Mn content in EPMA line scans of AlSi1.5Mn1 alloy cells grown
at growth rate 2 m/s. The measurement was taken at a marked line in figure above the
line scan.
In conclusion the line scan EPMA analyses reveal that significant purification in cells is
achieved. However, if an effective removal of the impurities from crystals is to be
achieved effective sweating process has to be performed in order to remove the impurity
rich intercellular material.
5.4.4. Inter-metallic compounds
Thermodynamic studies done for the alloying systems studied in this project, which
were carried out by our project partners at Corus and TNO [7], revealed that for ternary
alloys several intermetalic compound formations are possible. Intermetalic compounds
have an adverse effect on purification efficiency of the alloying systems. In this study

104

Aluminium scrap refining with solid layer fractional crystallization

several intermetalic compounds were found to be present in the intercellular regions of


the layer grown crystals.

Figure 5.10. Location of the measurements of the intermetalic compounds presented in


table 5.5.
Chemical compositions of the intermetalic compounds found in the intercellular region
of AlFe1.3Si0.3 alloying system are given in table 5.5. It is clearly seen that the only
intermetalic compound found for this alloying system is Al13Fe4 either pure ore mixed
with Al-fcc phase. The location of the samples where respectively 75 and 40 % Al13Fe4
was found to be present is shown in figure 5.10. Fluctuations from pure Al13Fe4
compound are due to the penetration of the x-ray during the measurement, which has
penetrated to the Al-fcc phase. In the intermetalic denoted as W in figure 5.10 the Al-fcc
phase is located below the Al13Fe4 phase, whereas in the intermetalic denoted as D in
figure 5.10 the Al13Fe4 is located under the Al-fcc phase. From thermodynamic
calculations was predicted formation of two other intermetalic compounds -AlFeSi and

-AlFeSi. However, from our extensive measurements they were not found to be present
on our samples. Formation of the intermetalic compound Al13Fe4 is the main cause for
decreasing the purification efficiency for iron for this alloying system and therefore
describes the reason of departure from thermodynamically predicted purification of iron
from this alloying system. In principle the intercellular impurities could be removed by

105

Chapter V

post purification processes such as sweating, which would have very positive effect on
increasing purification efficiency for this alloying system.
Table 5.5. The chemical composition of the intermetalic compounds found in the
intercellular region of AlFe1.3Si0.3 alloying system.
Al wt %

Si wt %

Fe wt % Compound

Comment

60.04

0.88

39.07

Al13Fe4

Calculated [7]

60.06

1.23

38.71

Al13Fe4

Measured in this work

69.99

0.77

29.24

75 % Al13Fe4 Intermetallic W in figure 9

83.65

0.58

15.51

40 % Al13Fe4 Intermetallic D in figure 9

Chemical compositions of the intermetalic compounds found in the intercellular space


of the AlCu4Mg1.4 alloying system and those calculated by Boender et al. are given in
table 6. As predicted by calculation there are two intermetalic compounds possible for
this alloying system. Both of them were found to be present in our samples. The AlCu-
intermetalic compound was found to be 90 % while due to the penetration of X-ray 10
% of Al-fcc phase appears in our measurements. Similar finding was the case for the
other intermnetalic compound. However AlCuMg-S intermetalic compound measured
in this work included 40 % of the Al-fcc phase.
Table 5.6. The chemical composition of the intermetalic compounds found in the
intercellular region of AlCu4Mg1.4 alloying system.
Al

Cu

Mg

Compound

Comment
Calculated [7]

38.1

44.8

17.1

AlCuMg-S

62.1

27.5

10.4

60 % AlCuMg-S Measured in this work

46.1

53.9

AlCu-

Calculated [7]

49.9

49.1

1.0

90 % AlCu-

Measured in this work

The intermetalics appearing in intercellular space of the AlSi1.5Mn1 were analysed and
results are presented in table 5.7. Based on calculations the presence of Al6Mn
intermetalic compound is expected and AlSi eutectic. However, Due to contamination
of the starting material the presence of Fe is measured, which complicates calculation of
the right percentage of intermetalics detected from our measurements.

106

Aluminium scrap refining with solid layer fractional crystallization

Table 5.7. The chemical composition of the intermetalic compounds found in the
intercellular region of AlCu4Mg1.4 alloying system.
Al

Si

Fe

Mn

Compound

Comment

58.63 10.57 5.90 24.90 Mixed compound Measured in this work


74.7

25.3

Al6Mn

Calculated [7]

5.5. Conclusions
Refining aluminium by means of layer growth fractional crystallization seems a
technically viable option in developing a sustainable technology for recycling of the
high-grade aluminium from scrap.
Chemical analyses reveal that significant purification is achieved. However, if an
effective removal of the alloying elements from aluminium scrap is to be achieved, a
repetition of the crystallization process has to be performed.
Microscopic examinations of the crystal samples, cut at different directions of crystal
growth, reveal that while crystals are growing as primary aluminium cells, an impurity
enriched melt is entrapped in the intercellular space.
Line scans reveal that concentration of impurities within cells is in line with
thermodynamic predictions while impurity rich area is detected in the intercellular
space.
In the intercellular space a number of intermetalic compounds were found to be present.
This leads to decreased purification efficiency of the layer grown crystals.
Application of the post-purification processes, such as washing or sweating, may
increase the purification efficiency significantly.

5.6. Acknowledgements
This research is carried out as a part of the research project Sustainable Technology for
the Reclamation of High-grade Aluminium from Scrap that is undertaken in the context
of the Dutch R&D stimulation programme on Economy, Ecology and Technology
(contract number EETK98026). For the experimental work, the assistance of Mr J.J.
Jansen of the Material Sciences Department is gratefully acknowledged.

107

Chapter V

5.7. References
1. Arkenbout G.J., 1995, Melt Crystallization Technology. Lancaster PA USA:
Technomic Publishing Company Inc.
2. Jansens P.J. & Matsuoka M., 2000, Melt Crystallization, in: Encyclopedia of
Separation Science, Eds. Wilson I.D., Adlard E.R., Cook M., and Poole C.F.,
Academic Press, San Diego, New York London, Sidney, Tokyo, 966-975.
3. Sillekens W.H., Verdoes D., and Schade van Westrum J.A.F.M., 1999,
"Refining Aluminium Scrap by means of Fractional Crystallization: Technical
Feasibility", Proceedings of the Fourth ASM International Conference and
Exhibition on the Recycling of Metals, ASM Europe, 105-114.
4. Esdaille J.D. and Walters G.W., 1975, "Lead Purification by Crystallization and
Reflux", Proceedings of the Aus. I.M.M. conference, South Australia, 187-195.
5. Pfann W.G., 1966, Zone Melting. John Wiley & Sons Inc., New York, NY,
U.S.A., 2nd edition.
6. Sillekens W.H., Schade van Westrum J.A.F.M., Bruinsma O.S.L., Mehmetaj B.,
Nienoord M.; 2000, "Refining Aluminium Scrap by means of Fractional
Crystallization: Basic Experimental Investigations"; Proceedings of the Fourth
International Symposium on Recycling of Metals and Engineered Materials, Ed.
D.L. Stewart Jr., J.C. Daley, and R.L. Stephens, Warrendale PA, USA, 963977.
7. Boender W., Waringa C.J., Krielaart G.P., Folkertsma A. and Verdoes D., 2002,
Refining Aluminium Scrap With Fractional Crystallization: Assessing Its
Feasibility With Thermodynamics, Proceedings of the TMS Annual Meeting &
Exhibition, Light Metals 2002. Ed.Schneider W., Seattle, Washington, USA,
TMS,1061-1068.
8. Sotome T. and Ohtaki M., 1998, "Application of Fractional Crystallization for
Refining of Molten Aluminium Scrap", Conference Proceedings of ICAA 6, vol.
1, Ed. T. Sato et al. Tokyo, Japan: The Japan Institute of Light Metals, 351-356.
9. Kahveci A.I. and Unal A., 2000, "Refining of a 5XXX Series Aluminum Alloy
Scrap by Alcoa Fractional Crystallization Process", Proceedings of the Fourth
International Symposium on Recycling of Metals and Engineered Materials, Ed.
D.L. Stewart Jr., J.C. Daley, and R.L. Stephens, Warrendale PA, USA: TMS,
979-991.

108

Aluminium scrap refining with solid layer fractional crystallization

10. P. Spencer, definition of thermochemical and thermophysical properties to


provide a database for the development of new light alloys, Proceedings of the
final workshop on COST 507, Volumes 1, 2 and 3, Vaals, The Netherlands 9-11
March 1997, EUR 18499, European communities, 1998.
11. J.T. Armstrong, Quantitative elemental analysis of individual microparticles
with electron beam instruments. In Electron Probe Quantitation, Heinrich,
K.F.J., and Newbury, D.E. (eds.), Plenum Press, (1991) 261-315.
12. M. Stepanski and E. Schfer, In Couirs on:Theory and Application of Melt
Crystallization, 28th 30th September 2000, Martin Luther University HalleWitwnbwrg, Halle (Saale), Germany.

109

Chapter V

110

CHAPTER VI
6. FRACTIONAL LAYER CRYSTALLIZATION AND SWEATING
APPLIED TO ALUMINIUM SCRAP REFINING
Abstract
In this chapter1 a fractional crystallization process for aluminium recycling based on
static layer growth will be described. Fractional layer crystallization is a separation
process by which the fractional separation is effected by directional crystallization from
the melt. As such it is an alternative to suspension crystallization. The main emphasis
will be paid to sweating as a post-purification process. Sweating is based on the partial
melting of the crystal layer by gentle heating close to the melting point of the pure
substance. The temperature during the sweating process is higher than in the
crystallization stage. The sweating process removes only the impurities entrapped in the
intercellular space, while impurities in the cells are not affected. Although during the
sweating process about 10 - 20 % of the crystals are removed, sweating is effective and
avoids an additional crystallization step in a fraction of the time.

Keywords: aluminium, layer crystallization, sweating, growth rate

Part of this work was published by Mehmetaj B., Kool W.H., Jansens P.J., and Katgerman L.:
Proceedings of the 7th International Conference on Semi-solid Processing of Alloys and Compounds (7th
S2P), Eds. Tsutsui Y., Kiuchi M., and Ichikawa K., National Institute of Advanced Sci. & Techn.- Japan
Society for Technology of Plasticity, Tsukuba, Japan, 24-28 September 2002, p. 611 - 616.

Chapter VI

6.1. Introduction
A common practice in aluminium recycling is the re-melting of aluminium scrap and
reuse for secondary alloys. In cases, where the alloying contents are higher than
according to the specification for a specific alloy, the molten scrap is diluted with
primary aluminium till the required chemical composition is achieved. With the
continuous increase of the aluminium consumption the amount of aluminium scrap that
needs to be recycled cannot be processed in this way in future. Therefore, closed loop
recycling is an imperative.
Melt crystallization is a separation process by which the fractional separation is effected
by directional crystallization from the melt. Fractional crystallization is applied for both
ultra purification of organic compounds [1,2] and metals [3]. Historically, the first
application of fractional crystallization for metals was the invention of the Pattison
process for extraction of silver from lead in 1829 [4]. It was followed by the
development of the zone melting technique in the 1950s by Pfann [5].

Fractional Crystallization

Suspension
Crystallization

Batch

Layer
Crystallization

Continuous

Static

Dynamic

Postpurification

Sweating

Diffusion
washing

Rinsing

Figure 6.1: Schematic representation of fractional crystallization.

Fractional crystallization is divided in two different approaches: suspension and layer


crystallization. A schematic representation is given in figure 6.1. The route studied in
this chapter is presented in bold. While in suspension growth crystals are grown as
small crystals suspended in the melt, in layer growth crystals are grown onto a cooled

112

Fractional layer crystallization and sweating applied to aluminium scrap refining

wall and at the end of the experiment are easily separated from the remaining liquid
(mother liquor), which is enriched with the impurities.
After growth of the layer sweating is used as a post-purification process. Sweating is
based on partial melting of the crystal layer by gentle heating close to the melting point
of the pure substance. Then, the impure melt adhered to the surface of the crystal layer
or present in the intercellular regions of the crystal layer drains off under the influence
of gravity. In this study both layer growth and the effect of sweating on the further
purification are described. To that purpose layers are grown onto a cold tube (cold
finger).

6.2 Experimental
A binary aluminium alloy with 6 wt % silicon is prepared by addition of an amount of
AlSi50 master alloy to the commercially pure aluminium melt. The ternary
AlCu4Mg1.5 alloy is prepared by adding of a pre-calculated amount of master alloys of
copper and magnesium to the commercially pure aluminium melt. The cold finger is
inserted in the melt and the temperature is lowered to the melting temperature of the
alloy by vigorous air blowing to the inside walls of the cold finger. The temperature is
stabilized for short period and then the cooling rate is adjusted by the airflow to achieve
the desired growth rate. The growth rate is determined by dividing the crystal layer
thickness by the growth time. After the crystallization stage the cold finger is removed
from the oven and is put in a preheated oven, which is used for the sweating stage.
Sweating is done at a constant temperature for about 10 15 minutes. The sweating
stage observed initially is called sweating A, and after some shaking of the cold finger is
called sweating B. Sweating B is carried out in order to remove the sweated liquid that
is entrapped due to oxide layer that is formed in the aluminium crystal layer during the
movement of cold finger with crystal layer from crystallization to sweating oven. The
weight of the sweated material is compared with the original weight of the grown layer
to obtain the percentage of the material lost as a result of sweating. The temperature
profile in the experiment for AlSi6 alloy is shown in figure 6.2. Similar curve pattern,
but with different temperature values, is true for AlCu4Mg1.4 alloy. The temperature
curve for the AlFe1.3Si0.3 and AlSi1.5Mn1 was much narrower. A more detailed
description of the experimental set-up is provided in [6] and is depicted in figure 6.3.

113

Chapter VI

Temperature/C

680
660

IV

640

II

III

620
600
0

0.5

1.5

2.5

3.5

4.5

Time/h

Figure 6.2. Temperature profile during the experiment with AlSi6: I: cooling interval;
II: layer growth; III: removal of the cold finger to preheated sweating oven; and IV:
sweating of crystals at constant temperature.

Temperature
controller

Sweating
oven

Crystallization
oven
Cold
finger

Figure 6.3. Equipment for Solid layer crystallization and sweating process.

The samples for chemical analyses were cut to include all the layers of the crystal. The
samples were dissolved in a mixture of 10 ml water, 10 ml HCl, and 6 ml HF. A closed
Teflon vessel was used to dissolve the samples in a microwave oven. Chemical analyses
of the alloys studied in this chapter were carried out using atomic absorption
spectroscopic technique (Perkin Elmer 2380(A)). Each sample was measured 3 times
and the results given in this chapter are average of these three measurements. The
overall error of chemical analyses is 3 %.

114

Fractional layer crystallization and sweating applied to aluminium scrap refining

Optical microscopy of the grown layers before and after sweating were taken in a
Neophot 30 inverted stage optical microscope equipped with a Polaroid digital camera
model DMC1 with DC input of 6 V, 3 A, connected to a PC.
The Electron Probe Micro Analaysis (EPMA) were performed with a JEOL JXA 8900R
microprobe using an electron beam with energy of 15 keV and beam current of 20 nA
employing Wavelength Dispersive Spectrometry (WDS). The composition at each
location of the sample was determined after background correction. The obtained
intensity ratios were processed with a matrix correction program CITZAF [7].

6.3. Results
In order to establish the potential of refining aluminium scrap by means of fractional
crystallization and sweating processes, AlSi6.2, AlFe1.3Si0.3, AlCu4Mg1.4, and
AlSi1.5Mn1 were selected. The first alloy was included as it represents a rather basic
and comprehensible model system. Other alloying systems were chosen as they
represent concentrations of the aluminium scrap streams in the Netherlands. The
obtained crystal layers of AlSi6.2 and AlCu4Mg1.4, before and after sweating were
analysed with chemical analyses, EPMA line scans, and their micrographic examination
was carried out. Sweating of the other two alloying systems; AlFe1.3Si0.3 and
AlSi1.5Mn1 was not successful. The main reasons for failure of sweating of these
alloying systems were: non ideal experimental conditions (sweating was carried out in
another oven and not in the same oven where crystallization took place), and the
temperature window for sweating of these alloying systems is to narrow.

6.3.1. Chemical analyses


Layer growth rate under the conditions applied was 2 m/s. Layers were grown to a
thickness of about 2 cm. Weight measurements showed that in sweating stage A, about
15 % and in sweating stage about B 23 % of the grown crystals were sweated off for
AlSi6.2 alloying system, while similar percentage were observed for sweating A of Al
Cu4Mg1.4 alloy. Sweating B of the later alloy was not successful and it ended up with
fallen crystal layer. This was due to a lesser stickiness of crystals to the cold finger of
the later alloy.

115

Chapter VI

The results of the chemical analyses of the layers grown before and after sweating for
AlSi6.2 alloying system are shown in table 6.1. Si concentration in the starting alloy
(C0) was 6.2 wt%. Si concentration in the grown layer (Ccr) is much lower and reduces
further after sweating. Si concentration in the sweated liquid (Csw) is high and even
higher than in the starting alloy. The overall distribution coefficient (koa = Ccr/C0) is also
given in table 6.1. It becomes lower during layer growth and sweating and it clearly
indicates that during layer growth purification has taken place and during the sweating
stage additional purification occurred.
Table 6.1: Chemical composition of the samples of ALSi6 alloy before layer growth,
after layer growth, and after two sweating stages A and B.
Starting

After

After

After

alloy

Layer Growth

Sweating A

Sweating B

C0 or Ccr (wt % Si)

6.20

3.90

3.40

3.08

Csw (wt % Si)

6.70

6.80

koa

0.63

0.55

0.50

Layer growth rate under the conditions applied for alloying system AlCu4Mg1.4 was 2
m/s. The results of the chemical analyses of the layers grown before and after sweating
are shown in table 6.2. The starting alloy was AlCu4Mg1.4. Although from mass
balance of starting master alloys it was expected that magnesium content of starting
alloy would be 1.51 per cent, due to its loss as a result of evaporation and dross
formation its content is significantly lower (see table 6.2). Copper and magnesium
concentrations in the grown layer (Ccr) are much lower and reduces further after
sweating, while their concentration in the sweated liquid (Csw) is high and even higher
than in the starting alloy. The overall distribution coefficient (koa) is also given in table
6.2. It becomes lower during layer growth and sweating and it clearly indicates that
during layer growth purification has taken place and during the sweating stage
additional purification occurred for this alloying system.
The impurity removal efficiency during the layer growth is significant for the applied
growth rate of 2 m/s (see table 6.2). For comparison of impurity removal efficiency
with theoretical possibilities for all alloying systems reader is referred to chapter 5. The
purification effect of the sweating process concerning copper is similar to that for AlSi6

116

Fractional layer crystallization and sweating applied to aluminium scrap refining

alloy. However, although during crystallization stage purification efficiency of Mg is


similar to Cu, after sweating Cu is removed with slightly higher efficiency. The
composition of the sweated liquid is not significantly higher than the original alloy and
is returned in the process.
Table 6.2. Chemical composition of the samples of AlCu4Mg1.4 alloy before layer
growth, after layer growth, and after two sweating stage.
Alloy

AlCu4Mg1.5

Element

Untreated

Crystals

After Sweating

C0/w%

Ccr/w%

Ccrsw (wt %)

Csw (wt %)

Cu

3.90

2.83

2.20

4.13

Mg

1.39

0.97

0.83

1.37

koa

Cu

0.72

0.56

koa

Mg

0.70

0.60

The purification efficiency of both crystal growth stage and sweating stage are relatively
high. However, if a further purification of the aluminium scrap has to be achieved lower
growth rates as well as repetition of the process have to be applied, in which the purified
layer is subjected to a subsequent process of layer growth and sweating, at a higher
temperature, which corresponds with the grade of the purification reached in the first
step.

6.3.2. Micrographs
In figure 6.4 cross-sections of the grown crystals of AlSi6 alloy before and after
sweating stages A and B are shown. These micrographs of the crystals clearly show that
with increasing amount of sweating the intercellular eutectic material is removed and
that the porosity of the crystals is increased significantly, i.e. the intercellular eutectic
material is removed.

117

Chapter VI

a)

b)

c)
Figure 6.4. Optical micrographs for the AlSi6 alloy from layer growth experiments: a)
before sweating; b) after sweating stage A; and c) after sweating stage B. The line in
the picture denotes 0.1 mm.

In figure 6.5 cross-sections of the grown crystals of AlCu4Mg1.4 before and after
sweating stage A are shown. These micrographs of the crystals clearly show that with
increasing amount of sweating the intercellular eutectic material is significantly
removed and that the porosity of the crystals is increased significantly.

118

Fractional layer crystallization and sweating applied to aluminium scrap refining

a)

b)

Figure 6.5. Optical micrographs for the AlCu4.4Mg1.4 alloy from layer growth
experiments: a) before sweating; b) after sweating stage A. The line in the picture
denotes 0.1 mm.

6.3.3. Line scans

The samples before and after sweating were analyzed using Electron Probe MicroAnalysis (EPMA) in order to determine the content of Si, Cu, and Mg in large grains
along a line and especially near the grain boundary. The points of analysis were located
along the line with increments of 4 m and involved the elements Al, Si, Cu, and Mg.

wt. % Si

AlSi6 Before Sweating

4
3
2
1
0
0

200

400

600

Position (m)
Figure 6.6. The Si content in EPMA line scan of AlSi6, grown at growth rate of 2 m/s,
before sweating, and position of the line scan (above).

119

Chapter VI

AlSi6 After Sweating

Si (wt- %)

3
2
1
0
0

200

400

600

800

1000

Position (m)

Figure 6.7. The Si content in EPMA line scan of AlSi6, grown at growth rate of 2 m/s,
after sweating, and position of line scan (above).
The Si composition profiles, given in figure 6.6, show that in a region of about 50 m
near the grain boundary the Si content gradually increases. In the middle of the large
grains the Si content is about 1 wt % and does not change after sweating (see figure
6.7). This shows that the sweating process has an effect on the removal of the
intercellular eutectic material, while at the same time the Si content in the grains is not
affected.
The Cu and Mg compositions profiles, given in figure 6.8, show that in the intercellular
region a highly impure material is present. The concentration of Cu is significantly
higher than that of Mg, which is shown on secondary axes. This explains the reason for
lower purification efficiency of Mg by applying sweating described in section 6.3.1.

120

Fractional layer crystallization and sweating applied to aluminium scrap refining

AlCu4Mg1.4 Intercellular space

wt %

Magnesium in secondary (Right) axes

10

60
50
40
30
20
10
0

8
Aluminium
Copper
Magnesium

6
4
2
0
1

Postion (m)

Figure 6.8. The Al, Cu and Mg content in EPMA line scans of AlCu4Mg1.4 alloy
intercellular material of crystals grown at growth rate of 2 m/s, and position of line
scan (above, line where white spots appear).

AlCu4Mg1.4 Sweated Sample

Copper

Magnesium

2,0

wt %

1,5
1,0
0,5
0,0
0

100

200

300

400

500

Position ( m )

Figure 6.9. The Cu and Mg content in EPMA line scans of AlCu4Mg1.4 alloy cells after
sweating, grown at growth rate 2 m/s, and position of line scan (above).
121

Chapter VI

In the middle of the large grains, as shown in figure 6.9, the Cu and Mg content is the
same with that reported in chapter V for the un-sweated sample, and does not change
after sweating. This shows that the sweating process has an effect on the removal of the
intercellular eutectic material, while at the same time the impurity content in the grains
is not affected. This is in line with the findings for AlSi6 alloy explained above.

6.4. Discussion

The impurity removal efficiency ( = 1 keff) for the AlSi6 alloy during the layer
growth is significant (37 %) for the applied growth rate of 2 m/s. The impurity
removal efficiency is significantly smaller than that thermodynamically possible (87 %).
The reason is that the tests were carried out at a relatively high growth rate. Microscopic
examination shows that while cells are grown impurities are enclosed in the intercellular
space as eutectic. The Si concentration in the cell interiors is close to that expected from
the phase diagram at the layer growth temperature (about 0.8 wt%). Similar results were
achieved also with the ternary alloy AlCu4.4Mg1.5.
Attempts to perform sweating to other two alloying system studied in this project,
AlFe1.6Si0.3 and AlMn1Si1.5 were not successful. The reasons are as follows: the
temperature frame of melting the crystal mass were very narrow and the temperature
control in the sweating oven was not accurate enough to perform a gentle heating in
order for the intercellular impurities to drain off. Hence, the intercellular impurities
were melted together with the pure crystals form cells. One has to mention that sweating
of these alloying systems is possible if better temperature control is applied, i.e.
sweating of crystals in the same oven where they grow.
From the micrographs of the studied alloys it is found that approximately 10 % of the
crystallized layer consists of eutectic intercellular material, which suggests that by
sweating maximum 10 % of the layer could be effectively removed. However, in
practice more material is removed by sweating. The reason is that besides entrapped
eutectic a significant amount of the crystals is melted, which opens the way for the
impurities to drain off, resulting in an amount of material sweated out, which is about

122

Fractional layer crystallization and sweating applied to aluminium scrap refining

twice as high as the amount of intercellular eutectic. However, due to some locked
impurity pockets not all the impurities could be removed by sweating process.
Sweating as a post-purification process reaches the same purity of crystals as was
observed during layer growth (without sweating) with 1 m/s growth rate [6].
Therefore, in order to obtain certain purification, one may grow a layer with the
required purity. The alternative strategy is to grow such a layer at faster growth rate,
which involves a higher impurity level, and to apply sweating which reduces the
impurity level to that required. It is found that the later strategy is better, since only half
of the time is needed. The composition of the sweated liquid is not significantly higher
than the original alloy and is returned in the process.
For an effective purification three strategies may be considered. The first is only layer
crystallization under static conditions; the second is a dynamic crystallization process,
which as a result of mixing improves mass and energy transfer increasing this way
purification efficiency; and the third is layer crystallization at a higher growth rat with
static or dynamic layer crystallization process, followed by a post-purification step such
as sweating. The later saves costs and energy.
The morphology of the crystal layers are affected by the applied growth rate and get
more stable with decreasing growth rate [6], resulting in a decreased amount of
intercellular eutectic material and an increased cell size, which leads to less liquid
pockets in the crystal layer and more open channels. As a result of this, it is anticipated
that by applying lower growth rates the effect of sweating in purification of the crystal
layer would increase and the amount of the crystal loss would decrease significantly. A
significant purification of the crystal layer is achieved in the process of layer growth and
sweating. However, if a further purification of the aluminium scrap has to be achieved
lower growth rates as well as repetition of the process have to be applied, in which the
purified layer is subjected to a subsequent process of layer growth and sweating, at a
higher temperature, which corresponds with the grade of the purification reached in the
first step.

6.5. Conclusions

Refining aluminium scrap by layer growth fractional crystallization is a viable


technology for recycling of aluminium scrap to high-grade secondary alloys. During

123

Chapter VI

layer growth purification has taken place and during the sweating stage additional
purification occurs.
Sweating process removes only impurities entrapped in the intercellular space, while
impurities entrapped inside the cells are not affected. With the increasing amount of
sweating the intercellular eutectic material is removed, while the impurity content
within the grains is not affected. The amount of material sweated out is higher than the
amount of the eutectic present, which means that also some crystals of the layer are
drained off.
For an effective purification three strategies may be considered. The first is only layer
crystallization under static conditions; the second is a dynamic crystallization process,
which as a result of mixing improves mass and energy transfer increasing this way
purification efficiency; and the third is layer crystallization at a higher growth rat with
static or dynamic layer crystallization process, followed by a post-purification step such
as sweating. The later saves costs and energy.

6.6. Acknowledgements
This research is carried out as part of the EET funded project Sustainable Technology
for the Reclamation of High-grade Aluminium from Scrap (contract number
EETK98026). Fruitful discussions at the Technical Project Group Meetings with the
EET partners and in particular with Dr. D. Verdoes (TNO-MEP, Apeldoorn) and Dr. P.
de Vries (Corus RD&T, IJmuiden) are gratefully appreciated. For the experimental
work, the assistance of Mr J.J. Jansen of the Material Sciences Department is gratefully
acknowledged.

124

Fractional layer crystallization and sweating applied to aluminium scrap refining

6.7. References

1. Arkenbout G.J.: Melt Crystallization Technology. Lancaster PA USA:


Technomic Publishing Company Inc., 1995.
2. Jansens P.J. and Matsuoka M.: Melt Crystallization, in: Encyclopedia of
Separation Science, Eds. Wilson I.D., Adlard E.R., Cook M., and Poole C.F.,
Academic Press, San Diego, New York London, Sidney, Tokyo, (2000) 966975.
3. Drini B., Katgerman L., and Boom R.: Metal Refining with fractional
crystallization: State-of-the-art and future prospects, Accepted at Steel
Research International.
4. Esdaille J.D. and Walters G.W.: "Lead Purification by Crystallization and
Reflux", Proceedings of the Aus. I.M.M. Conference, South Australia, (1975)
187-195.
5. Pfann W.G.: Zone Melting. John Wiley & Sons Inc., New York, NY, U.S.A.,
2nd edition,1966.
6. Mehmetaj B., Bruinsma O.S.L., Kool W.H., Jansens P.J., and Katgerman L.:
Aluminium scrap recycling with solid layer fractional crystallization, 15th
International Symposium on Industrial Crystallization (ISIC-15), Sorrento, Italy,
1518 September 2002, p.879 884.
7.

Armstrong J.T.: Quantitative elemental analysis of individual microparticles


with electron beam instruments. In Electron Probe Quantitation, Heinrich,
K.F.J., and Newbury, D.E. (eds.), Plenum Press, (1991) 261-315.

125

Chapter VI

126

CHAPTER VII

7. EVALUATION OF THE ALUMINIUM FALLING LIQUID


FILM CRYSTALLIZATION PROCESS STEPS

7.1. Introduction
Falling film crystallization process proved to have several advantages over the static
crystallizer for organic compounds. As a result of the flowing film the liquid becomes
turbulent. The turbulence has positive effect on enhancing both energy and mass
transfer of the crystallizer. Due to this effect falling film crystallizer for organic
compounds achieves the same purification efficiency, as for static crystallizer, with four
to five times higher growth rate [1].
Although there are considerable differences between organic compounds and metals
behaviour during fractional crystallization the basic processes are similar. The main
problems for adopting this process for metals is high operation temperatures and
corrosiveness of metal melt to equipment. However, mass and heat transfer of metals
are higher and makes production rates much faster.
In this chapter an evaluation of the falling film process for aluminium scrap refining as
well as equipment design and building will be described. This will be done to see the
potential of following research in this direction in the future.

7.2. Hydrodynamics of the thin falling liquid films


The flow of liquids in thin films is a common observation in everyday life; such as flow
of rainwater in roads, window panes, and roofs. The flow of liquid films is frequently
used in different industrial applications. In this report, the case of thin liquid film flow
on the walls of a heat exchanger for fractional crystallization of aluminium scrap will be
analyzed in more details.
A dimensional analysis of the flow of liquid films shows that the regime of flow
depends on the value of the dimensionless Reynolds number, Re, which is defined as;

Chapter VII

Re =

- is the density of the falling liquid film;


- is the film thickness;
- is dynamic viscosity of the falling liquid film; and
v - is the mean flow velocity of the liquid film.

As it is seen in the expression above, the Reynolds number expresses the ratio between
the inertia to the viscous forces.
Based on the value of the Re the regime of flow in thin films is divided as follows:

1. Smooth laminar flow

0 < Re < Re1

2. Wavy laminar flow

Re1 < Re < Re2

3. Turbulent flow

Re2 < Re

Different authors use different values of the Reynolds number to describe the transfer
from one regime of flow to the other. For review of the flow of liquids in thin films,
where a table of authors and critical Re numbers for transition from laminar to turbulent
flow regime is given by Fulford [2]. However, generally accepted values for transition
from one flow regime to another are: Re1 from 2 for falling films in vertical wall to 10
for inclined walls; Re2 is considered to be between 250 and 500, with the Re of 400
widely used in falling film fractional crystallization processes.
According to Brauer [3] the mean thickness of a falling laminar liquid film is expressed
in the form;

3 2
=
g sin

1/ 3

Re1 / 3

- is kinematic viscosity (/);


g- is gravity constant; and
- is the slope of the wall in which the falling liquid film flows.

128

Evaluation of the Aluminium Falling Liquid Film Crystallization process steps

Film thickness in mm

Falling film Thickness


0.5
0.45
0.4
0.35
0.3
0.25
0.2
0.15
0.1
0

200

400

600

800

1000

Re

Figure 7.1. The thickness of the aluminium falling liquid film on the vertical plate, as a

function of Re number, for laminar flow regime.

The film thickness of the aluminium falling film in vertical plate depending on Re
number for laminar flow regime is plotted in figure 7.1.

7.3. Heat transfer in falling liquid film


The heat transfer in falling liquid films is expressed as a function of dimensionless
Nusselt number, which is a function of Reynolds (Re) and Prandtl (Pr) numbers in
problems of forced convection (Nu = f(Re, Pr)) . The Prandtl number is expressed as a
ratio of kinematic viscosity, , to thermal diffusivity, , (Pr = /). The value of the Pr
number for liquid metals is 0.0010.04. Due to the high thermal conductivity of metals
the thermal diffusion layer, T, is much higher than the falling film thickness (T ).
For liquid metals, with a uniform heat flux at the wall as a boundary condition, the
result for heat transfer of a laminar falling film in a flat plate are approximated as [4]

0.88

Nu = Re Pr *

1
+
1
.
317
Pr

129

Chapter VII

7.4. The effect of Reynolds number in diffusion boundary layer


thickness
Although turbulence on the falling film appears around Re = 400, a thin boundary layer
exists close to the wall where the transport appears by atomic diffusion only. This
boundary layer becomes turbulent only at Re value of 5-10 105 [5]. The diffusion
boundary layer thickness, D, can be expressed in the form [6]:

D
*
D =

Nu * m *

0.344

m*- is the molar ratio between impurity and the mixture


D- is impurity diffusion coefficient

This expression shows that the diffusion layer thickness decreases proportionally with
the film thickness , which in tern decreases with decreasing kinematic viscosity and
increasing the slope of the falling film. Figure 4 shows the diffusion boundary layer as
a function of Reynolds number for vertical liquid aluminium falling film. The diffusion
coefficient of most of the liquid metals is within the range of 1-10 * 10-9 m2/s.
Therefore a value of 5 * 10-9 was taken as a representative for the metals present in
aluminium alloys that will be studied in this project, and was used in calculations for
figure 7.2.

diff. layer thickness in


mm

diff. layer
0,035
0,03
0,025
0,02
0,015
0,01
0,005
0

200

400

600

800

1000

Re

Figure 7.2. The diffusion boundary layer of an aluminium falling liquid film over a

vertical wall as a function of Reynolds number.


130

Evaluation of the Aluminium Falling Liquid Film Crystallization process steps

7.5. Optimization of the aluminium falling liquid film crystallizer


Based on the theoretical description made in previous sections of this report we will try
to predict the optimum conditions for purification of aluminium alloys, studied in this
project, using film crystallization process. The main criteria to be considered here are:
growth rate, optimal crystal production rate, optimal feed temperature, and optimal feed
rate.

7.5.1. Optimal growth rate.

The optimum criteria for crystal growth rate can be found from the BPS-model
described in the chapter III of this thesis:

k eff =

k0
k 0 + (1 k 0 ) exp G *

( )

and

G* =

G c
*
kd l

keff- is effective distribution coefficient;


k0- is thermodynamic distribution coefficient;
G*- is the dimensionless crystal growth rate;
kd- is coefficient of the mass transfer rate ;
G- is crystal growth rate; and

c and l- are respectively the densities of solid and liquid.

131

Chapter VII

This criterion describes the minimization of the ratio of contaminant concentration in


crystal to that in residue. According to Arkenbaut [7] the optimum criteria for effective
removal of the impurities is when G * 0.1 , and as explained in the second biannual

report, depends on the value of the thermodynamic distribution coefficient.

7.5.2. Optimal crystal production rate.

The optimal dimensionless crystal growth rate determines the crystal production rate.

mc = c * Gopt * A = c * G * *

*A

Where,

mc- is the crystal production rate; and


A- is the surface area of the crystallizer.

This expression shows that the decrease of the diffusion layer thickness, D, increases
crystal productivity. Diffusion layer thickness, D, decreases with increasing mass feed
rate (see figure 4). Therefore an optimum feed mass rate that decreases diffusion layer
thickness, D, is important for the design of the falling film crystallizer.

7.5.3. Optimal feed temperature.

The temperature of the aluminium feed should be high enough to prevent interface
stability. In the same time a very high feed temperature causes problems for the uniform
layer thickness. For organic systems a feed temperature of 5-10 K above the melting
point of the feed system is found to be optimal [Mayer, 1973].

132

Evaluation of the Aluminium Falling Liquid Film Crystallization process steps

7.5.4. Optimum feed rate (Re).

The feed rate determines the Reynolds number defined as:

Re =

mf
w l

Where,

mf- is mass feed rate (kg/s); and


w- is the width of the falling film.

At low feed rates the interface stability of the crystal surface is not stable. For organic
compounds a minimum Re = 50 is as a criteria to avoid this effect [7, 8]. As shown in
figure 4 the diffusion layer thickness decreases with increasing Re. However, at Re >
1000 the decrease is negligible. Hence, the reasonable upper value of Reynolds number
is 1000.

7.5.5. Optimum crystal growth rate for aluminium falling film process.

For this purpose the following data will be taken into account:

D = 5 * 10-9 m2/s
Re = 1000
G* = 0.1

D = 1. 34 * 10-5 m (derived from figure 7.2)

According to Ossipov, the optimum growth rate can be calculated as follows:

133

Chapter VII

Gopt = G *

= 0 .1 *

5 *10 9 m 2 s 1
= 3.73 *10 5 ms 1
5
1.34 *10

From this approximate calculation can be seen that the optimum crystal growth rate in
an aluminium falling film crystallization process can be in the range of 10-5 m/s, which
is 10 times as fast as for most of the organic systems.

7.6. Design of the aluminium falling film crystallizer


As described in previous sections of this chapter, by generating a falling film a
significant improvement of the heat and mass transfer can be achieved. Following
favourable theoretical predictions, development of the dynamic (falling film) layer
crystallization process for aluminium scrap refining is a feasible thing to do. This
process is patented by TU Delft [9, 10]. A basic design proposal for the falling film
process is presented in figure 7.3.
The falling film crystallizer contains a top reservoir with an opening in the bottom that
can be controlled to let the outflow of the small amounts of aluminium to the cooling
plate. The cooling plait is 5 cm wide and 30 cm long. It is situated in a tube oven, which
is heated from the front to keep the plat close to the melting point of the crystallizing
alloy. The plait is cooled from the backside by cold air to remove the heat of fusion that
is released during the crystal layer growth. The aluminium melt flowing from the upper
reservoir is superheated to prevent solidification in the opening. It flows as a thin film
on the cooled plait, where only a small fraction of the melt is crystallized and the rest
falls to the bottom reservoir. The bottom reservoir is used to keep the melt from
solidifying and can be used for the next crystallization step, preventing this way energy
loss that would be needed to re-melt the alloy. Cooling power of the called finger
controls layer growth rate.

134

Evaluation of the Aluminium Falling Liquid Film Crystallization process steps

Al melt
at Tm + 5 K

Cooling
Plate

10 dm3/h
aluminium

5 cm
Falling
Al thin film
Tm

30 cm

Residue

Figure 7.3. Schematic drawing of the aluminium falling film crystallizer.

Aluminium melt flow rate versus Reynolds number for the newly designed falling film
crystallizer is presented in figure 7.4. For an optimum enhancement of the mass transfer
a typical flow rates with Re = 1000 is considered to be an optimal in melt crystallization
world. It clearly shows that flows of 10 litters per hour would be optimal for the newly
designed aluminium falling film crystallizer.

135

Chapter VII

Al flow
12
liters/hour

10
8
6
4
2
0
0

200

400

600

800

1000

Re

Figure 7.4. Aluminium melt flow rate versus Reynolds number for the newly designed

falling film crystallizer.

7.7. Equipment for falling film crystallization process


The aluminium melt flowing from the upper reservoir is superheated to prevent
solidification in the opening. However, as first experimental trials showed, superheating
alone is not enough to prevent that. That was the reason for building new top reservoir
which has heater closer to the bottom to provide more heating close to the valve. The
same problems were accounted, although in a smaller scale, with the new reservoir.
After initial flow through the valve aluminium solidifies and blocks further flow. To
overcome this problem new heaters around the valve are installed. The aluminium melt
flows as a thin film on the cooled plate, where only a small fraction of the melt is
crystallized and the rest falls to the bottom reservoir. The bottom reservoir is used to
keep the melt from solidifying and can be used for the next crystallization step,
preventing this way energy loss that would be needed to re-melt the alloy. Cooling
power of the cold finger controls layer growth rate.
Due to time restrains this equipment will be used in an follow-up project and no results
from crystallization will be presented in this thesis.

136

Evaluation of the Aluminium Falling Liquid Film Crystallization process steps

Figure 7.5. Technical drawing of the aluminium falling film crystallizer.

7.7.1. Control of the crystallizer


The control scheme of the aluminium falling film crystallizer is shown in figure 7.6. All
heating is done by resistance heaters. The control of the temperature as well as pressure
in the crystallizer is done by computer using Labview program. The control is based on
the Pulse Width Modulation principle. In this a complete pulse has a width of 2 seconds.
When the PI controller has an output of 40%, the heating elements will receive power
for 0.8 seconds, every two seconds. The power switching is done by solid state relays.

137

Chapter VII

At the moment the molten aluminium flow is in contact with the cooling plate the
heating of the plate is replaced by controlled cooling. This cooling is done by
compressed dry air, flowing through small pipes and than spread by a filter element of
sinter metal. The cooling air control valve is proportionally controlled from the
computer. On the top of the top-reservoir extra pressure is applied to provide constant
flow of aluminium melt through the valve.

Figure 7.6. The control scheme of the aluminium falling film crystallizer

138

Evaluation of the Aluminium Falling Liquid Film Crystallization process steps

7.8. Conclusions and recommendations


Falling film crystallization process proved to have several advantages over the static
crystallizer. As a result of the falling film the liquid becomes turbulent. The turbulence
has positive effect on enhancing both energy and mass transfer of the crystallizer. Due
to this effect falling film crystallizer for organic compounds achieves the same
purification efficiency, as for static crystallizer, with four to five times higher growth
rate [6].
Theoretical analysis of the falling film process for aluminium scrap refining reveal that
the optimum growth rate of such crystallizer is within the order of 10-5 m/s, which is
about 10 times faster than optimum growth rate for static process. This increase in the
optimum growth rate is attributed to the improved mass and heat transfer due to the
turbulence that appears as a result of the melt flow down the heat exchanger wall. The
higher growth rates compared to static crystallizer make this process suitable for a
process that requires high production rates, which with aluminium scrap recycling is the
case.
For an optimum enhancement of the mass transfer a typical flow rates with Re = 1000 is
considered to be an optimal in melt crystallization world. It clearly shows that flows of
10 litters per hour would be optimal for the built aluminium falling film crystallizer.

139

Chapter VII

7.9. References
1. J. Ulrich, in Theory and Application of melt Crystallization, 28th 30th
September 2000, Martin Luther University Halle Witwnbwrg, Halle, Germany
2. G.D. Fulford, in Advances in Chemical Engineering, Volume 5, Academic
Press, New York 1964.
3. H. Brauer, VDI Forschungsheft, 457 (1956) 22.
4. D.R. Poirier and G.H. Geiger, Transport phenomena in materials processing,
TMS, Warrendale, 1994.
5. H. Schlichting, Boundary-layer theory, McGraw-Hill, New York, 1955.
6. Ossipov, Appl. Math. Modelling, 23 (1999) 419.
7. G.J. Arkenbout, Melt Crystallization Technology, Technomic Publishing
Company, Lancaster, PA, U.S.A., 1995.
8. M.U. Mayer, ber die gerichtete fraktionierte Kristallisation aus dem
Rieselfilm, Ph.D Thesis, ETH Zrich, 1973.
9. B. Drini, W.H. Kool, P.J. Jansens, and L. Katgerman: Method of Recovering a
Metal from a Mixture, World patent no. PCT/ WO/2004/085686, 07 of October
2004.
10. B. Drini, W.H. Kool, P.J. Jansens, and L. Katgerman: Werkwijze voor het
winnen van een metaal uit een mengsel, patent no. NL1023009C, 1 December
2004.

140

ALUMINIUM SCRAP REFINING WITH FRACTIONAL LAYER


CRYSTALLISATION
SUMMARY
Primary aluminium production is an energy intensive process. With an increase of the
aluminium use in car manufacturing a recycling technological process is imperative. The
energy use for recycling is only a fraction of the energy needed to produce primary
aluminium. To this end a joint project with Corus plc in IJmuiden, TNO-MEP in Apeldoorn
and TNO Institute of Industrial Technology in Eindhoven was carried out as a part of the
research project Sustainable Technology for the Reclamation of High-grade Aluminium
from Scrap, which is undertaken in the context of the Dutch R&D stimulation programme
on Economy, Ecology and Technology (contract number EETK98026). While our partners
worked on a suspension based fractional crystallisation process, our research at Delft
University of Technology was concentrated on layer based fractional crystallization
technique for aluminium scrap purification.
Historically speaking, three stages can be recognized in the development of fractional
crystallization processes for the purpose of metal refining. A first milestone was the
invention of the Pattinson process for the extraction of silver from lead in 1833, the second
was the development of the zone-melting technique in 1952, and the third one is the socalled Yunnan crystalliser for refining of tin as of 1975, which is probably the most
advanced application for metals at the moment. Fractional crystallization, widely known in
the metals world as fractional solidification, is a separation technique applied for
purification of metals and organic melts. While most of the currently available processes
are mainly developed for the production of high-purity metals (99.99 wt % and more), the
emerging technologies appear to be exploring the possibilities for aluminium scrap
recycling. For the latter application, proposed methods have not left the laboratory stage yet
with research in this field still ongoing. General restrictions are that the processes are quite
slow and of limited production capacity. Also, the different requirements on the formation
of the crystals (concerning crystal growth rates and stirring) and the separation of the solid

141

Summary

from the liquid fraction are in most cases not fully met. Consequently, there still is room for
further progress.
Fractional crystallization technique is much more advanced for organic compounds than for
metals. This is mainly due to the better process conditions, such as low melting temperature
and lower reactor attack, of organic materials. However, a careful study of possibilities of
adopting these processes for metals will open new horizons. Prospects are good for both
modes of operation: suspension and layer crystallization. The most important metals that
could use this technique are aluminium, silicon, magnesium, lead, and precious metals.
Zone melting is a relatively simple and useful experimental technique to assess the ability
to refine distinct alloying constituents from complicated alloy systems (like industrial kinds
of scrap aluminium). Regarding the obtained purification in the experiments, results
basically comply with the estimates based on the binary alloy phase diagrams, although
there are known distinct deviations from other experimentally achieved results. Chemical
analyses reveal that purity of the samples on the head and on the middle positions are
almost the same, which reveals that purification takes place in more than the half of the
original samples.
Microscopic examination of the AlSi6 alloy reveals that while aluminium cells grow, the
impurities that rich eutectic concentration are trapped between them. The AlSi0.4Fe1.6
alloying system micrographs show a flat interface till about 40 per cent of the sample is
crystallized and later a eutectic growth resumes. The other two systems studied in this
thesis (AlCu4Mg1.5 and AlSi1.5Mn1.1) grow with flat interface and no change in their
morphology was observed by microscopy.
In chapter 5 a fractional crystallization process for aluminium recycling based on static
layer growth is described. Static layer crystallization experiments showed that for AlSi7.2
alloying systems, typical growth rates of 1 to 2 m/s are optimal for aluminium scrap
purification. In order to achieve the preset purification criteria for the pure product, one or
more re-crystallizations steps are required. Depending on the impurity concentration, the
crystallization temperature varies from 600 to 660 C.

142

Aluminium scrap refining with fractional layer crystallisation

Refining aluminium by means of layer growth fractional crystallisation is technically a


viable option in developing a sustainable technology for recycling of a high-grade
aluminium from scrap. Chemical analyses reveal that significant purification is achieved.
However, if an effective removal of the alloying elements from aluminium scrap is to be
achieved a repetition of the crystallisation process has to be performed. Microscopic
examinations of the crystal samples, cut at different directions of crystal growth, reveal that
while crystals are growing as primary aluminium cells, an impurity enriched melt is
entrapped in the intercellular space. Line scans reveal that concentration of impurities
within cells is in line with thermodynamic predictions while impurity rich area is detected
in the intercellular space. In the intercellular space a number of intermetalic compounds
were found to be present. This leads to decreased purification efficiency of the layer grown
crystals.
Application of the post-purification processes, such as washing or sweating, may increase
the purification efficiency significantly. During layer growth purification has taken place
and during the sweating stage additional purification occurs. With the increasing amount of
sweating the intercellular eutectic material is partially removed, while the impurities within
the grains are not affected. Sweating is based on the partial melting of the crystal layer by
gentle heating close to the melting point of the pure substance. The temperature during the
sweating process is higher than in the crystallization stage. Although during the sweating
process about 10 - 20 % of the crystals are removed, sweating is effective and avoids an
additional crystallization step in a fraction of the time. The amount of material sweated out
is higher than the amount of the eutectic present, which means that also some crystals of the
layer are drained off. For an effective purification two strategies may be considered. The
first is only layer crystallization and the second is layer crystallization at a higher growth
rate, followed by a post-purification step such as sweating. The later saves costs and energy.
Falling film crystallization process proved to have several advantages over the static
crystallizer. As a result of the flowing film the liquid becomes turbulent. The turbulence
has positive effect on enhancing both energy and mass transfer of the crystallizer. Due to

143

Summary

this effect falling film crystallizer for organic compounds achieves the same purification
efficiency, as for static crystallizer, with four to five times higher growth rate.
Theoretical analysis of the falling film process for aluminium scrap refining reveals that the
optimum growth rate of such crystallizer is within the order of 10-5 m/s, which is about 10
times faster than optimum growth rate for static process. This increase in the optimum
growth rate is attributed to the improved mass and heat transfer due to the turbulence that
appears as a result of the melt flow down the heat exchanger wall. The higher growth rates
compared to static crystallizer make this process suitable for a process that requires high
production rates, which with aluminium scrap recycling is the case. Further (experimental)
studies are needed, whereby several technical challenges associated with falling film
crystallization of aluminium are addressed, before the techno-economic potential of this
new technology can be established unambiguously.

144

ALUMINIUM SCHROOT OPWERKING MET FRACTIONERENDE LAAG


KRISTALLISATIE

SAMENVATTING
Primaire aluminiumproductie is een energie-intensief proces. Met een toename van
het

gebruik

van

aluminium

in

de

autoproductie

is

een

technologisch

terugwinningsproces noodzakelijk. De energie die nodig is voor terugwinning is


slechts een fractie van de energie welke nodig is voor de primaire productie. Daarom
is een gezamenlijk project met Corus plc in IJmuiden, TNO MEP in Apeldoorn en het
Instituut van Industrile Technologie in Eindhoven uitgevoerd als onderdeel van het
onderzoeksproject

Duurzame

Technologie

voor

het

Terugwinnen

van

Hoogwaardig Aluminium uit schroot, dat is ondernomen in het kader van het
Nederlandse

Onderzoeksstimulatieprogramma

voor

Economie,

Ecologie

en

Technologie (contractnummer EET98026). Terwijl onze partners werkten aan een


suspensie-gebaseerd fractionerende kristallisatie proces, was ons onderzoek aan de
Technische Universiteit Delft geconcentreerd op de laag-groei fractionerende
kristallisatie techniek voor de opwerking van aluminium schroot.
Historisch gezien kunnen er drie fasen onderscheiden worden in de ontwikkeling van
fractionerende kristallisatie processen voor metaalopwerking. Een eerste mijlpaal was
de uitvinding van het Pattinsonproces voor de extractie van zilver uit lood in 1833, de
tweede was de ontwikkeling van de zonesmelt techniek in 1952 en de derde is de
zogeheten Yunnan kristallisator in 1975. De laatste is waarschijnlijk de meest
geavanceerde toepassing voor metalen op dit moment. Fractionerende kristallisatie,
algemeen bekend in de metaalwereld als fractional solidification, is een
scheidingstechniek toegepast voor zuivering van metalen en organische smelten.
Terwijl de meeste momenteel beschikbare processen vooral ontwikkeld zijn voor de
productie van hoogzuivere metalen (99.99 gewichtsprocent en meer), lijken de
opkomende technologien de mogelijkheden te onderzoeken om aluminium schroot te
hergebruiken. Voor de laatste toepassing hebben de voorgestelde methoden de
laboratoriumfase nog niet achter zich gelaten, maar het onderzoek op dit gebied duurt
nog steeds voort. Algemene beperkingen zijn dat het proces nogal langzaam is en dat

Samenvatting
de productiecapaciteit beperkt is. Ook wordt in de meeste gevallen nog niet volledig
voldaan aan de verschillende voorwaarden aan de vorming van de kristallen (wat
betreft de kristalgroeisnelheid en het roeren) en de scheiding van de vaste- en de
vloeistoffractie. Als gevolg hiervan is er nog ruimte voor verdere vooruitgang.
De fractionerende kristallisatie techniek is veel geavanceerder voor organische stoffen
dan voor metalen. Dit is vooral het gevolg van de voordeliger procescondities, zoals
het lagere smeltpunt en de mindere reactorcorrosie, van organische materialen. Toch
zal een nauwgezet onderzoek naar de mogelijkheden om deze processen te gebruiken
voor metalen, nieuwe horizonten openen. De vooruitzichten zijn goed voor de beide
procesmethoden: suspensie- en laagkristallisatie. De belangrijkste metalen waarvoor
deze techniek gebruikt zou kunnen worden zijn aluminium, silicium, magnesium, lood
en kostbare metalen.
Zonesmelten is een relatief eenvoudige en bruikbare experimentele techniek om de
mogelijkheid vast te stellen om bepaalde legeringsbestanddelen te raffineren uit
gecompliceerde legeringen (zoals industrile aluminium afvalsoorten). Wat betreft de
behaalde zuiverheid in de experimenten komen de resultaten ongeveer overeen met de
schattingen gebaseerd op binaire fasendiagrammen van legeringen, hoewel er
bepaalde afwijkingen zijn van andere experimenteel behaalde resultaten. Chemische
analyses tonen aan dat de zuiverheden van de monsters aan de kop en in het midden
bijna hetzelfde zijn, hetgeen aantoont dat zuivering plaats vindt in meer dan de helft
van de originele monsters.
Microscopisch onderzoek van de AlSi6 legering toont aan dat tijdens de groei van de
individuele aluminium cellen ertussen vloeistof wordt ingebouwd die verrijkt is qua
onzuiverheden. De microscoopfotos van de AlSi0.4Fe1.6 legering laten een plat
oppervlak zien totdat ongeveer 40 procent van het monster gekristalliseerd is, daarna
treedt eutectische groei op. De andere twee systemen die bestudeerd zijn in deze
dissertatie (AlCu4Mg1.5 en AlSi1.5Mn1.1) groeien met een plat oppervlak en er werd
geen verandering in morfologie geobserveerd met de microscoop.
In hoofdstuk 5 is een een fractionerende kristallisatie proces beschreven voor
aluminium terugwinning gebaseerd op statische laag-groei. Statische laag146

Aluminium schroot opwerking met fractionereende laagkristallisatie


kristallisatie experimenten lieten zien dat, voor AlSi7.2 legeringen, typische
groeisnelheden van 1 tot 2 m optimaal zijn voor aluminium schroot opwerking. Om
de

vereiste

productzuiverheden

te

behalen

zijn

of

meerdere

herkristallisatiestappen nodig. Afhankelijk van de concentratie van de onzuiverheden


varieert de kristallisatietemperatuur van 600 tot 660C.
Opwerking van aluminium door middel van laag-groei fractionerende kristallisatie is
technisch een haalbare methode om een duurzame technologie te ontwikkelen voor
het hergebruik van hoogwaardig aluminium uit schroot. Chemische analyses tonen
aan dat een significante opzuivering behaald wordt. Niettemin moet, om een
effectieve verwijdering van de legeringselementen te bewerkstelligen, het
kristallisatieproces meermaals uitgevoerd worden. Microscopisch onderzoek van de
kristalmonsters, doorgesneden in verschillende kristalgroeirichtingen, toont aan dat er,
terwijl de kristallen groeien als primaire aluminiumcellen, een qua onzuiverheden
verrijkte smelt ingebouwd wordt in de intercellulaire ruimte. In de intercellulaire
ruimte werd de aanwezigheid van een aantal intermetallische stoffen geconstateerd.
Dit leidt tot een lagere zuiveringsefficintie van de laag-gegroeide kristallen.
Toepassing van post-zuiveringsprocessen zoals zweten, kan de zuiveringsefficintie
significant

doen

toenemen.

Gedurende

de

laag-groei

heeft

opzuivering

plaatsgevonden en gedurende de zweetfase vindt er additionele zuivering plaats.


Tijdens het zweten wordt het het intercellulaire eutectische materiaal gedeeltelijk
verwijderd, terwijl de onzuiverheden in de eigenlijke aluminium cellen niet worden
benvloed. Zweten is gebaseerd op het gedeeltelijk smelten van de kristallaag door
voorzichtig te verwarmen tot vlakbij het smeltpunt van de zuivere stof. De
temperatuur gedurende het zweten is hoger dan in de kristallisatiefase. Ondanks het
feit dat bij het zweten 10-20 % van de kristallen wordt verwijderd, is zweten effectief
en voorkomt het een additionele kristallisatiestap in een fractie van de tijd. De
hoeveelheid uitgezweet materiaal is groter dan de aanwezige eutectische hoeveelheid,
hetgeen betekent dat ook enkele kristallen van de laag worden verwijderd. Voor een
effectieve opzuivering kunnen er twee strategien in overweging genomen worden.
De eerste is alleen laag-kristallisatie en de tweede is laag-kristallisatie bij een hogere
groeisnelheid, gevolgd door een post-zuiveringsstap zoals zweten. Het laatste bespaart
kosten en energie.
147

Samenvatting

Het vallende-film kristallisatieproces bleek meerdere voordelen te hebben ten


opzichte van de statische kristallisator. Als gevolg van de stromende film wordt de
vloeistof turbulent. De turbulentie heeft een positief effect door het verbeteren van de
energie- en stofoverdracht in de kristallisator. Hierdoor behaalt een vallende-film
kristallisator eenzelfde zuiveringsefficintie als een statische kristallisator, met een
vier tot vijf keer hogere groeisnelheid.
Theoretische analyse van het vallende-film proces voor aluminium schroot opwerking
toont aan dat de optimale groeisnelheid voor zon kristallisator in de orde van 10-5 m/s
is, hetgeen ongeveer 10 maal groter is dan de optimale groeisnelheid voor het
statische proces. Deze grotere optimale snelheid wordt toegeschreven aan de
verbeterde massa- en warmteoverdracht als gevolg van de turbulentie die optreedt
wanneer de smelt langs de wand van de warmtewisselaar stroomt. De hogere optimale
groeisnelheid ten opzichte van het statische proces maakt het proces geschikt voor
processen waarbij hoge productiecapaciteiten vereist zijn, hetgeen het geval is bij de
opwerking van aluminium schroot. Verder (experimenteel) onderzoek is nodig,
waarbij verschillende technische uitdagingen, geassocieerd met vallende-film
kristallisatie van aluminium, onderzocht worden, voordat de technisch-economische
potentie van deze nieuwe technologie met zekerheid kan worden vastgesteld.

Bedri DRINI

148

RAFINIMI I SKRAPIT T ALUMINIT ME KRISTALIZIM FRAKSIONAL N


SHTRES

PRMBLEDHJE
Prodhimi i aluminit primar sht nj proces q krkon shum energji. Me rritjen e sasis s
prdorimit t aluminit pr prodhimin e automjeteve, zhvillimi i nj procesi teknologjik pr
riciklimin e aluminit sht i domosdoshm. Energjia e nevojshme pr riciklimin e aluminit
sht vetm nj pjes e vogl e energjis q do t duhej pr ta prodhuar aluminin nga xehja.
Pr t kontribuar n zhvillimin e nj procesi t riciklimit t aluminit nj projekt i prbashkt
i hulumtues i Universitetit Teknik t Delftit me partnerve Corus plc nga IJmuideni, TNOMEP nga Apeldorni dhe TNO Instituti i Teknologjis Industriale nga Eindhoven sht
zhvilluar si pjes e projektit hulumtues Teknologji e Qndrueshme pr Riciklimin e
Aluminit nga Mbeturinat , I cili sht ndrmarr n kuadr t programit hulumtues
holandez, i cili stimulon programet pr Ekonomi, Ekologji dhe Teknologji (kontrata me
numr EETK98026). Prderisa partneret ton e kan hulumtuar procesin e kristalizimit
fraksional

n suspenzion, grupi yn hulumtues n Universitetin Teknik t Delftit ka

hulumtuar procesin e kristalizimit fraksional n shtres pr riciklimin e aluminit.


Kur flasim nga aspekti historik i zhvillimit t procesit t kristalizimit fraksional pr
rafinimin e metaleve, mund ti identifikojm tri faza. Zbulimi i par ka qen procesi i
Patisonit pr nxjerrjen e argjendit nga przierja e tij me plumb n vitin 1833, i dyti sht
zbulimi i teknikes s shkrirjes zonale nga Pfani n vitin 1952, dhe i treti sht i
ashtuquajturi procesi Junan rafinimin e kallajit n vitin 1975, i cili konsiderohet si procesi
me i avancuar pr rafinimin e metaleve pr momentin. Kristalizimi fraksional sht teknik
e ndarjes q prdoret pr rafinimin (pastrimin) e metaleve dhe kompozimeve organike.
Prderisa shumica e proceseve aktuale t kristalizimit fraksional aplikohen pr prfitimin e
metaleve me pastrti t lart 99.99 % e me shum), teknologjit q po hulumtohen jan
prqendruar n zhvillimin e proceseve teknologjike pr rafinimin e mbetjeve metalike t
aluminit. Aplikimet e fundit jan ende n fazn laboratorike dhe hulumtimet shkencore n
kt fushe jan duke vazhduar. Kufizimet kryesore t ktyre proceseve jan se ato jan t

Prmbledhje

ngadalshme dhe me kapacitete t kufizuara. Gjithashtu, krkesat e prgjithshme pr


formimin e kristaleve (shpejtsia e kristalizimit dhe przierja gjat kristalizimit) si dhe
ndarja e kristaleve nga przierja e tyre me material t lngt, n shumicn e rasteve nuk i
plotsojn krkesat teknologjike. Prandaj, n kt fush ka mjaft mundsi pr prparim n
kt lmi..
Teknika e kristalizimit fraksional sht shum m e prparuar pr kompozimet organike se
sa pr metale. Arsyeja kryesore pr kt sht q proceset pr materie organike zhvillohen
n kushte me t favorshme si nga aspekti i temperaturs ashtu edhe nga aspekti i
korrozionit t pajisjeve t reaktorve n t cilt kryhet kristalizimi. Megjithat, nj studim i
kujdesshm i mundsis s adaptimit t proceseve qe prdoren pr kompozime organike n
rafinimin e metaleve do t hapte horizonte t reja n fushn e kristalizimit fraksional.
Parashikimet jan t mira si pr kristalizimin n shtres ashtu edhe pr kristalizimin n
suspenzion. Metalet me t rndsishm q do t mund t prfitonin nga zhvillimi i ktyre
teknikave jan alumini, silici, magnezi, plumbi dhe metalet e muar
Shkrirja n zon sht metod relativisht e thjesht eksperimentale pr t vlersuar aftsin
rafinimit t legurave t ndryshme t komplikuara, si jan mbeturinat e ndryshme t skrapit
t aluminit. Sa i prket efektin e pastrimit t aluminit nga legurat e tij, rezultatet jan n
linj m pritjet pr kto legura q vijn nga diagramet e fazave pr legura me binare, edhe
pse ka devijime t pritura nga rezultate eksperimentale. Analizat kimike tregojn s n
shumicn e rasteve pastrtia e aluminit sht e njjt si n fillim ashtu edhe n mes t
mostrave t kristalizuara me shkrirje zonale, q do t thot se pastrimi i legurave t
studiuara t aluminit ndodhe n me tepr se gjysma e gjatsis s mostrs.
Analiza me mikroskop e mostrave t kristalizuara (mikrografi) t legurs AlSi6 tregojn s
derisa kristalet n form gishtash t aluminit rriten, papastrtit t cilat arrijn prqendrimin
eutektik grumbullohen n hapsirn n mes tyre. Mikrografet e legurs AlSi0.4Fe1.6
tregojn nj siprfaqe t rrafsht derisa t kristalizohet prafrsisht 40 % e mostrs dhe
pastaj vazhdon me rritje eutektike t kristaleve. Legurat tjera t studiuara me metodn e
shkrirjes zonale (AlCu4Mg1.5 dhe AlSi1.5Mn1.1) rriten me siprfaqe t rrafsht dhe nuk
mund t vrehen ndryshime me ndihmn e mikroskopit.

150

Rafinimi i skrapit t aluminit me kristalizim fraksional n shtres

N kapitullin 5 prshkruhet procesi statik i kristalizimit fraksional n shtres pr riciklimin


e aluminit. Eksperimentet e kristalizimit statik treguan se pr legurn AlSi7.2 shkalla e
rritjes se shtress prej 1 deri n 2 m/s sht optimale. Nse pastrtia e parashikuar duhet t
arrihet ather prsritja e kristalizimit sht e nevojshme. Varsisht nga prmbajtja e
papastrtive n skrap temperatura e kristalizimit arrin nga 600 deri n 660 C.
Rafinimi i mbeturinave t aluminit me metodn statike t kristalizimit fraksional sht
teknikisht i

mundshm. Analizat kimike tregojn se me kt teknik pastrim i

konsiderueshm i legurave t aluminit mund t arrihet. Megjithat nse pastrtia e krkuar


duhet t arrihet ather pr t gjitha legurat e studiuara n kt studim prsritja e procesit
t kristalizimit sht e domosdoshme. Analizat me mikroskop t mostrave t kristalizuara
me kt metode jan t ngjashme me ato t kristalizuara me shkrirjen n zon. Skanimet n
linja tregojn se koncentrimi i papastrtive n kristale sht n linj me parashikimet
termodinamike, kurse n hapsirn n mes kristaleve jan grumbulluar pastrtit t cilat
ngurtsohen me koncentrik

eutektik. Kjo shkakton ulje t efikasitetit t pastrimit t

kristaleve.
Aplikimi i teknikave t post-pastrimit, si jan djersitja e kristaleve dhe larja e tyre, e rrisin
n mnyr t konsiderueshme. Gjat kristalizimit pastrimi i aluminit ndodh kurse gjat
djersitjes se kristaleve ato pastrohen edhe me shum. Me rritjen e sasis se kristalit t
djersitur rritet edhe pastrtia e tyre. Me rritjen e sasis s djersitur t shtress rritet sasia e
mbetjes eutektike qe sht e vendosur n mes t kristaleve primare, kurse papastrtit q
gjinden n kristalet primare nuk largohen. Djersitja bazohet n shkrirjen parciale t shtress
kristalore duke e ngrohur ngadal deri t pika e shkrirjes s kristaleve. Temperatura gjat
djersitjes sht m e lart sesa gjat kristalizimit. Edhe pse gjat djersitjes 10 20 % e
kristaleve humbin, procesi i djersitjes sht efektiv dhe mnjanon nj kristalizim shtes t
nevojshm pr pastrimin e skrapit n nj kohe shumfish m t shkurt. Sasia e mass
kristalore t humbur gjat djersitjes sht me e madhe sesa sasia e mbetjes eutektike q
gjendet n mes t kristaleve primare, q do t thot se edhe nj pjes e kristaleve primare
shkrihet. Pr nj pastrim efikas mund t konsiderohen dy mnyra. Njra sht t rrisim
shtresn kristalore ngadal, kurse tjetra t rrisim at me shpejt dhe pastaj ta djersitim at.
Metoda e dyt kursen koh dhe para.

151

Prmbledhje

Kristalizimi n metodn e filmit rrshqits ka shum prparsi n krahasim m metodn


statike t kristalizimit n shtres. Si rezultat i lvizjes s filmit rrshqits likuidi behet
turbulent. Turbulenca ka efekt pozitiv si n transferin e mass ashtu edhe n transferin e
energjis n kristalizator. Si rezultat i ktij efekti kristalizatort e till pr materie organike
mbrrijn t njjtin efekt t pastrimit sikurse kristalizatori statik me shpejtsi kristalizimi 4
deri n 5 her me t madhe.
Analizat teorike t procesit t filmit rrshqits pr pastrimin e skrapit t aluminit tregojn se
shpejtsia e rritjes se shtress kristalore sht rreth 10-5 m/s, q sht rreth 10 here me
shpejt sesa me procesin statik. Kjo rritje e efikasitetit i atribuohet rritjes se transferim t
mass dhe energjis si rezultat i lvizjes s shkrirjes metalike n muret kmbyesit e
nxehtsis. Shpejtsia e rritjes s shtress kristalore t ktij procesi n krahasim m
procesin statik, e bn procesin dinamik (filmi rrshqits) m t prshtatshm pr procese qe
krkojn kapacitet prodhuese me t mdha, q n rastin e skrapit t aluminit sht e
domosdoshme. Para s t vrtetohet potenciali tekniko ekonomik i ksaj metode, studime t
mtutjeshme eksperimentale jan t nevojshme pr t tejkaluar nj numr t
konsiderueshm t sfidave teknike qe kan t bjn me metodn e filmit rrshqits.

Bedri DRINI

152

ACKNOWLEDGEMENTS

I would like to express my gratitude to number of people who have contributed to the
completion of this book.
First of all I would like to thank Prof. Ir. Laurens Katgerman and Prof. Dr. ing. Dolf
Bruinsma for giving me the opportunity to carry out this research at TU Delft.
Special thanks go to my promoters Prof. Ir. Laurens Katgerman and prof. Dr. ir. Peter
Jansens for their guidance and support during my research. Last year my stay outside the
Netherlands has made too difficult to complete this work. Due to Prf. Dr. ir. Peter Jansenss
persistence and patience, and his assistants Leslie van Leeuwens help in completing all
the administrative procedures, it was possible to finalize this work.

I am grateful to Prof. Dr. ing. Dolf Bruinsma for his support and guidance during the first
year of my research work. He taught me the fundamentals of fractional crystallization and
was always ready to discuss all the problems I was facing during that initial period of my
work.
I am very grateful to the project technical group people: Dr. ir. Dirk Verdoes and Ir.
Michil Niewenoord of TNO-MEP in Apeldoorn; Dr. ir. Wim Boender and Dr. ir. Paul de
Vries from Corus Ijmuiden, Dr. ir. Wim Sillekens, jan Schade van Westrum and Mark van
Haut form TNO Industry. Their support during the TPG meetings was very useful and in
the same time it helped me understand the issues of suspension based process.
I am also grateful to people with whom I have discussed issues considered in this thesis and
with whom I have written couple of articles: Prof. Dr. ir. Rob Boom, Dr. Ir Pim Kool and
others.
I am thankful to my colleagues and friends from my group: Dr. Dmitri Eskin, Dr. Lidy
Appachitei-Fratila, Dr. Iulian Appachitei, Dr. Jan Zuiderma, Dr. Volodia Bilovol, Suyitno,
and others for making a friendly atmosphere during coffee brakes and their support with all
scientific and personal consultations during my stay in the Netherlands. I am also grateful
to the help given by researchers at the API department for useful discussions during
monthly meetings.

Acknowledgments
My experimental work would not be completed successfully if there were not the talented
group of GST technical group: Jan van Etten, Tjerd Tobi, Jan-Peter Boomsma. Special
thanks go to Jaqcue Jansen for his mastership on dealing with molten aluminium and great
help with my experimental setup.
Last but not least my thanks go to my family and friends for inspiration and support during
my research at TU Delft. I thank my wonderful wife, Marika, for her selfless support and
encouragement. My daughters Teut and Liza were my endless source of joy and
inspiration.

154

PUBLICATIONS
Publicistic articles:
1. Bedri Mehmetaj; Hazards from polluted water, (in Albanian), Daily Rilindja,
Prishtin - Zrich - Tiran, 10 June 1996.
2. Bedri Mehmetaj; QSL- A new method of extraction of lead, (in Albanian),
Daily Rilindja, Prishtin Zrich - Frankfurt-Tiran, 20 Jun 1998.
3. Bedri Drini; Air pollution shortens your live, (in Albanian), Daily Bota Sot,
Prishtin - Zrich, 27 Oktober 2002.
Reports and Studies:
1. Verdoes, D, Nienoord, M, Boender, W, Landskroon, JPS, Vries, PA de,
Drini, B

, Hout, MHFM van & Sillekens, WH (2002). Hoogwaardig

aluminium uit schroot, Final Report, 2004. Apeldoorn, IJmuiden: TNO,


Corus,

28

pp.

http://www.senternovem.nl/mmfiles/EETK98026_tcm24-

166126.pdf.
2. Drini, B. and Dushi, M. Study for prospects of metal industry development
in Kosovo, financed by the Ministry of Trade and Industry, NKI Drini
Consulting and Engineering, Prishtina, Kosova, 2005, pp 52.
Journal and Conference Papers
1. Sillekens W.H., Schade van Westrum J.A.F.M., Bruinsma O.S.L., Mehmetaj B.,
Nienoord M.; "Refining

Aluminium Scrap

by

means

of

Fractional

Crystallisation: Basic Experimental Investigations"; Proceedings of the Fourth


International Symposium on Recycling of Metals and Engineered Materials,
Eds. Stewart Jr. D.L., Daley J.C., and Stephens R.L., TMS, Warrendale PA,
USA (2000): p. 963977.
2. Bedri Mehmetaj and Rob Boom: Ferronickel production in Kosovo Past
performance and new opportunities, steel research 72 (2001) No. 11-12, p.
428-433.
3. Mehmetaj, B , Haasnoot, JG, Cola, L de, Albada, GA van, Mutikainen, I,
Turpeinen, U & Reedijk, J: Syntheses, Characterization, X-Ray Crystal
155

Publications
Structure, Redox and Photophysical properties of Polypyridyl Ruthenium(II)
Complexes Containing Carboxylate Substituted Pyridyltriazoles; European
Journal of Inorganic Chemistry, (2002), p. 1765 1771
4. Mehmetaj B., O.S.L. Bruinsma, W.H. Kool, P.J. Jansens, L. Katgerman
Aluminium scrap recycling with solid layer fractional crystallization,
Proceedings of the 15th International Symposium on Industrial Crystallisation
(ISIC-15),

Ed.

Chianese,

A.,

Sorrento,

Italy,

Chemical

Engineering

Transactions, Volume 1, 2002, p.879 884.


5. Mehmetaj B., W.H. Kool, P.J. Jansens, L. Katgerman, Fractional layer
crystallization and sweating applied to aluminium scrap recycling, Proceedings
of the 7th International Conference on Semi-solid Processing of Alloys and
Compounds (7th S2P), Eds. Tsutsui Y., Kiuchi M., and Ichikawa K., National
Institute of Advanced Sci. & Techn.- Japan Society for Technology of Plasticity,
Tsukuba, Japan, 24-28 September 2002, p. 611 - 616.
6. Drini B., Katgerman L., and Boom R.: Metal Refining with fractional
crystallization: State-of-the-art and future prospects, Proceedings of the ECI
Conference on Metal Separation Technologies III, June 20-24 2004, Copper
Mountain, Colorado, USA, Eds.

Aune, R.E. and Keukkonen M., Helsinki

University of Technology, Laboratory of Metallurgy, Espoo, Finland; p. 34-41..


7. Drini B., Katgerman L., and Boom R.: Metal Refining with fractional
crystallization: State-of-the-art and future prospects, Accepted in Steel
Research International.
Patents:
1. Drini at al.: Method of Recovering a Metal from a Mixture, (World Patent)
PCT/ WO/2004/085686, 07 of October 2004. In web address:
http://www.wipo.int/pctdb/en/wo.jsp?KEY=04/85686.041202
2. Drini at al.: Werkwijze voor het winnen van een metaal uit een mengsel,
(European Patent) NL1023009C, 1 December 2004. In web address:
http://v3.espacenet.com/textdoc?DB=EPODOC&IDX=NL1023009C&F=0&QP
N=NL1023009C

156

CURRICULUM VITAE

CURRICULUM VITAE
Name:

Bedri DRINI (former MEHMETAJ)

E-mail:

bedri.drini@dce-group.com

Date of birth:

February. 7. 1970

Education:

1984 1988 Gymnasium in Peja, Kosova


1988 1989 Military Service in Zagreb, Croatia
1989 1996 University of Prishtina, Faculty of Natural Sciences
and Mathematics, Department of Chemistry (BSc)
1996 1997 Central European University, Department of
Environmental Sciences and Policy, Budapest,
Hungary (MSc.).
1997 1999 doctoral researcher in Chemistry Department of
the Leiden University, The Netherlands.
1999 present Ph.D. Researcher at Delft University of
Technology, Delft, The Netherlands.

Work Experience:

1997 1999

Other activities:

Guest Researcher in Chemistry Department of


the Leiden University, The Netherlands
(concluded with Drs.).
1999 present Doctoral researcher at Delft University of
Technology, Delft, The Netherlands.
2004
-Taught an intensive course on Environmental
Sciences at the University of Prishtina sponsored
by Brain Gain Program of World University
Services, Austrian branch (WUS-Austria).
2005 present Co-owner of DCE- Drini Consulting &
Engineering Company, based in Prishtina.
2006
Kosova National development plan- Office of
Primeminister, Coordinator for sectorial strategies
for following sectors: Energy, Mining, Transport,
Education, and High value-added services.
2006 President of the Board of Business Consultants
Council (BCC), Association of consulting
companies in Kosova.

157

You might also like