Sei sulla pagina 1di 113

Lab course Crystal Growth I, Master of Science Crystalline Materials

Experiment: Float-Zone Growth of Silicon




Relevant Literature:

Attached scripts

J. Bohm, A. Ldge and W. Schrder: Crystal growth by floating zone melting. In: Handbook
of Crystal Growth 2a, Ed. D.T.J. Hurle (Elsevier/North-Holland, Amsterdam 1994), 213

K.Th. Wilke and J. Bohm: Kristallzchtung. (VEB Deutscher Verlag der Wissenschaften,
Berlin 1988, and under license, Harri Deutsch, Thun 1988).

W. Keller and A. Mhlbauer: Floating-zone silicon. Vol. 5 of: Preparations and properties of
solid-state materials, Ed. W.R. Wilcox (Marcel Dekker, NewYork 1981)

W. G. Pfann: Zone melting. 2nd Ed.(J. Wiley, New York 1966)



Task:

FZ growth of an 8-10mm diameter silicon crystal in a double ellipsoid mirror furnace
(MHF).

Step 1: Introduction to the furnace and equipment. Observation of an FZ experiment done by
the advisor (1
st
day)

Step 2: Setup of seed and feed rod in the furnace, FZ growth of the crystal (2
nd
day)

Experiment protocol: the protocol should contain a brief introduction to the process as well as
relevant data (material used, power used, growth rate etc.) and a short discussion of the
results.

Material used: doped (P or As or Sb) silicon, orientation (111) or (100).

Dangers involved: The process involves high-voltage electricity, vacuum and pressurized
process chambers and high temperatures. No changes to the equipment are allowed without
consent of the advisor.










MHF description (in
German)






2. The floating-zone process
The general setup of the floating-
zone process is shown in fig. 2-1: A
small free melt (or solution) volume,
held only by surface tension and
adhesion, is suspended between the
growing single crystal and a
polycrystalline feed rod. Under earth
conditions, hydrostatic pressure due
to gravity causes the characteristic
bottle shape of the zone. Crystal
growth is achieved by a relative
movement of the crystal and feed rod
versus the melt zone, i.e. the heater.
2.1 History and practical considerations
The FZ process can be regarded as a special variant of the zone
melting process invented by Kapitza [Kap28] for the crystal growth of
bismuth, and later developed by Pfann [Pfa52, Pfa66] with respect to the
purification (zone refining) and doping (e.g. zone leveling) of semicon-
ductor material. Zone melting employs a container for the material to be
processed, usually a long tube, boat or ampoule, and is used in horizontal
and vertical configurations. A small portion of the material is kept
molten by a suitable heater which is then moved relative to the material.
For impurities or dopants with segregation coefficients k=c
s
/c
l
1
(where c
s
and c
l
are the concentrations in the solid and liquid,
2. The floating-zone process 5
Fig. 2-1: Schematic drawing of
the floating-zone process.
respectively) a redistribution profile characteristic for the process is
obtained (fig. 2-2, see chapter 3 for details) with a concentration reduc-
tion in the first part of the material and a corresponding increase in the
last zone for segregation coefficients k < 1 and vice versa for coefficients
k > 1. The impurity reduction can be further improved by several zone
passes (fig. 2-2), including using several heaters at a time. In addition to
the purification effect, advantage can be taken of the plateau region of
the initial profile to achieve doping profiles of constant concentration.
For the same reason, incongruently melting materials can be grown from
the melt by this process, because after the initial transient the zone has
the peritectic composition in equilibrium with the solid.
6 2. The floating-zone process
Fig. 2-2: Calculated (eqs. 3-I and 3-II) axial dopant segregation curves
for a Si:P crystal (c
o
= 610
17
cm
-3
, k
o
= 0.35) grown by zone melting with a
zone length of 12mm, for 1 zone pass and 3 zone passes, respectively.
In the floating-zone process, the container is omitted and the melt
zone is suspended between the growing crystal and the feed material as
shown in figs. 2-1, 2-3 and 2-4. The floating-zone process was first
described and introduced in the fifties, by Theuerer [The52], Keck and
Golay [Kec53], and Emeis [Eme54]. Its first application was the crystal
growth of silicon [Kec53] and this remains the dominant industrial appli-
cation of the process to this day [Zuh89, Boh94]. Nevertheless, within
the last 40 years a large variety of materials has been grown by this
method, ranging from semiconductors (fig. 2-3) to refractory metals,
oxides (fig. 2-4), halides and others (see [Boh94], pages 244, 245,
247-249 for a comprehensive listing).
2. The floating-zone process 7
Fig. 2-3: Silicon (mp: 1410
o
C)
floating zone of 10mm in a
double ellipsoid mirror furnace
(see section 2.1.1.1b).
Fig. 2-4: Gadolinium Gallium
Garnet (Gd
3
Ga
5
O
12
, GGG, mp:
1767
o
C) floating zone of 4mm
in a double ellipsoid mirror
furnace, from [Ger84]
Variations of the process are: a) the standing pedestal method, where
the feed rod, located at the bottom, is much larger than the crystal above
it, and the crystal is pulled from a pool of melt at the top of the large rod.
This method, introduced by Dash [Das58, Das60] and Poplawsky and
Thomas [Pop60], can be regarded as a cross between the CZ and the FZ
technique and is often used for the production of crystalline optical fibers
(see e.g. [Fei86, Ima95]); a hanging pedestal method, with the positions
of feed rod and crystal reversed, is also possible. b) the ribbon to ribbon
technique (RTR) invented by Lesk [Les76] and Gurtler [Gur78] for the
recrystallization of silicon sheets for photovoltaic applications, uses a
very asymmetric zone, where the height and the thickness of the zone are
orders of magnitude smaller than the width. This geometry poses some
specific problems with regard to dimensional control, interface shape and
heat transfer [Yec95]. In addition to ribbons, the production of crystal
tubes by the FZ technique has also been studied [Pfa66, Gle89, Lan94a,
Lan94b, Hsi96].
In principle, it is also possible to substitute the floating melt zone of
the FZ process by a floating solution zone, resulting in the FSZ (or
traveling solvent floating zone, TSFZ) method. It can be regarded as a
derivation of the traveling heater method (THM, see e.g. [Ben79,
Ben80]) invented by Broder and Wolff [Bro63], similar to the way the
FZ process was derived from zone melting. The process appears advan-
tageous with respect to crystal quality, because the lower growth
temperature leads to a defect reduction, and due to the absence of the
ampoule wall the stress in the peripheral parts of the crystal is reduced.
That this effect is indeed possible has been shown by growth experi-
ments with GaSb crystals from free Ga solutions [Ben80, Ben82].
A free zone of near peritectic composition may be used for growing
incongruently melting materials with aggressive fluxes, such as YFe
2
O
4
,
8 2. The floating-zone process
Y
3
Fe
5
O
12
(YIG) [Kim77, Kim78, Kit79, Shi79, Fei88], or high T
c
super-
conductors [Fei88, Gaz88, Gaz89]. Other materials grown from
(partially) free solvent zones are CaCO
3
from a Li
2
CO
3
flux [Bri71,
Bel72, Bel76], Ba
0.65
Sr
0.35
TiO
3
from a TiO
2
flux [Hen74], and LaB
6
from
La or B fluxes [Ota92, Ota93]. Due to the usually slow growth rates of
solution growth compared to melt growth (mm/day vs. mm/h - mm/min),
the FSZ process has been used rather seldom.
Due to the action of gravity on the liquid, the majority of experiments
is done in a vertical configuration, although horizontal systems employ-
ing electromagnetic levitation to counteract gravity have been used
[Pfa56, Pfa66]. For the same reason, the pulling direction of the crystal is
usually parallel to the gravity vector under earth conditions (compare
also section 2.2). Exceptions are the standing pedestal method and
processes with very small zones (e.g. RTR) where the influence of
gravity on the zone shape is not as important.
The actual growth equipment differs widely depending on the heater
concepts (section 2.1.1) and the materials used. For zone translation, it is
possible to move either the heater or the crystal and feed rod. The first
solution is preferable in terms of disturbances and mechanical vibrations,
but the latter allows two independent translation and rotation mecha-
nisms to achieve a much better control of the zone and interface shapes.
Especially important is the fact that with two translation drives it is
possible to employ the necking process first introduced by Dash [Das59]
and Ziegler [Zie61] for the growth of dislocation-free silicon and germa-
nium. In this technology, also used in Si-CZ growth, a thin "neck" of
material is produced by employing a larger translation rate of the crystal
compared to the feed rod. Dislocations introduced by the seed or the
thermal shock upon melting propagate to the surface of the crystal and
2. The floating-zone process 9
disappear
a
; after achieving dislocation-free growth, the crystal diameter
is increased to its final value. As the critical resolved shear stress neces-
sary to form new dislocations is much larger than the stress necessary to
multiply existing ones, the crystal remains free of dislocations. This
process is mainly possible for elemental semiconductors (Si and Ge),
because their critical resolved shear stress for the formation of new dislo-
cations is much higher than that of most compound semiconductors. For
heavy crystals such as in the industrial FZ-Si production, the thin neck
makes the attachment of an additional supporting mechanism above the
neck necessary [Boh94].
As starting material for FZ growth a compact (and preferably cylindri-
cal) rod is necessary; this can either be obtained from solid (poly)crystal-
line material prepared by other melt growth processes, deposited from
the gas phase, or be pressed and/or sintered from powder. In the latter
case the remaining porosity of the rod has to be taken into account in the
mass balance, i.e. the pulling speeds of crystal and feed rod. Additional
problems can arise through the formation of bubbles released from a
porous feed rod [Ger90] or an absorption of zone material through capil-
lary action by a feed rod with open porosity. An additional run for
compacting the material is helpful in these cases, see e.g. [Pfa66, Hig95].
Employing an oriented seed crystal to achieve single crystal growth is
the most common method, but it is possible to make a seed selection by
necking if no single crystal is available.
For the translation/rotation mechanisms, avoiding mechanical vibra-
tions is of paramount importance; floating zones are very effective vibra-
tion sensors. Likewise, the resonance frequencies of the growth
equipment should be far from those of the zone (the latter are usually in
10 2. The floating-zone process
a
an exception are orientations where the directions of growth and dislocation
movement coincide, such as [110] in the diamond lattice.
the range of 0.1 to 10 Hz, depending on the size and material) and the
crystal. Otherwise, resonance in the system can lead to a disruption of a
zone near its stability limit, or even break the crystal in the necking
region [Boh94].
Control signals used for automation include image analysis of video
pictures, scanning of the zone and crystal shapes by lasers, absorption of
radiation from a radioactive source [ Aut75, Lub86], or analysis of the
torque when employing differential rotation [Que75, Lub86]. Weighing
the crystal and/or the feed rod, as in some Czochralski systems, is not
very common. Manual control and observation of the zone, especially in
the critical phases of seeding, necking, and increasing the diameter is
still widely employed.
2. The floating-zone process 11
2.1.1 Heater concepts
The fact that only a small volume of the material to be processed is
heated to temperatures above melting point (mp), and that usually no
other material must be heated, allows for a much larger than usual
variety of furnace types for the FZ process. A rough classification leads
to the main groups of radiation heating (fig. 2-5), high frequency heating
(fig. 2-21), electron beam and plasma heating (fig. 2-24), and direct
heating (figs. 2-26, 2-27).
12 2. The floating-zone process
Fig. 2-5: Radiation heating methods: (a) electrical resistance heating,
(b) heating by focused light, employing either an ellipsoidal mirror
(top) or two parabolic mirrors (bottom), (c) laser heating (adapted
from [Car91]); E is a conventional beam expander and A an axicon-
based ring beam expander.
2.1.1.1 Radiation heating
Radiation heating includes the classic resistance heater concept,
heating by focused light (image/mirror furnaces), and laser heating. Fig.
2-5 illustrates the principles of these methods.
2.1.1.1a Resistance heating
Resistance heating furnaces for floating-zone processes can be
anything from a simple ringwire as in fig. 2-5a to elaborate versions with
several central heaters, cooling rings, isothermal heaters and after heaters
as in the so-called zone melting facility (ZMF) shown in fig. 2-6. The
more elaborate furnaces allow the individual tailoring of the temperature
profile for a given material system, e.g. an optimization of the axial
temperature gradient with respect to growth conditions at the interface
and with respect to thermal stress in the grown crystal region. Most resis-
tance heating elements are made either from Kanthal (operating tempera-
tures up to 1200
o
C) or from PtRh alloys (usual operating temperatures up
to 1400
o
C). For higher temperatures, resistance heating is more difficult,
possible heating elements being made from Pt/Ir, MoSi
2
(Super Kanthal),
SiC, or carbon (graphite, CFC, or carbon on BN), the latter requiring an
inert gas atmosphere or vacuum for operation.
A major drawback of most resistance heaters with operating tempera-
tures above approximately 800
o
C is that visual control of the melt zone,
either directly or via some optical and/or electronic system, is nearly
impossible or leads to a significant disturbance of the thermal field. This
problem complicates crystal growth considerably, as the knowledge of
the zone shape is vital for the control of the process (see section 2.2).
Under such circumstances, one has to rely on temperature information
2. The floating-zone process 13
from the heating elements, experience and/or numerical simulations of
the heat transfer in the system for optimizing control parameters. Heater
control itself is easily achieved with thermocouples and the usual PID
controllers; for high precision requirements (relative accuracies of
14 2. The floating-zone process
Fig. 2-6: Concept of the zone melting furnace ZMF developed by the
aerospace company Dornier, after [Beh89, Len90, Sch92].
0.01K), optical fiber thermometry of the heating elements can be used
[Dol95].

2.1.1.1b Imaging furnaces
Imaging furnaces using focused light are a rather special type of heater
with a distinct set of advantages and limitations. They can be very
energy efficient, there is no principal limit to processing temperatures,
and the visual control of the zone and the crystal growth process is excel-
lent. On the other hand, temperature measurements during growth are
practically impossible. The temperature in the heating elements (i.e. the
lamps) is only indirectly related to the sample temperature, and contact-
less measurement of the sample temperature by pyrometry is also nearly
impossible due to the much higher level of light reflected versus radia-
tion emitted from the sample [Eye77, Eye81]. Information on the
temperature field can only be obtained from special measuring samples
with incorporated temperature sensors, and from numerical simulations.
Starting in the fifties and sixties [Wei56, Bau59, Koo61], different
imaging furnaces have been developed over the years for floating-zone
applications, e.g. [Fie68, Aka69, Tri70, Sau71, Cox72, Ars73a, Ars73b,
Miz74, Kit77, Eye77, Eye79, Eye81, Bal81, Bed84, Car84, Car86a,
Len90, Mat92, Bal93], some of them for microgravity experiments on
manned and unmanned space flights (compare figs. 2-7, 2-8, 2-10 and
2-16). The original rationale for preferring image furnaces for that appli-
cation was the absence of electromagnetic interference in comparison to
radio frequency heating (section 2.1.1.2) and, in comparison to resistance
heating, the small weight and volume in relation to the maximum
temperature obtainable. At the moment, commercial systems specifically
designed for floating-zone growth are available from three different
2. The floating-zone process 15
16 2. The floating-zone process
Fig. 2-7: Monoellipsoid mirror furnace ELLI developed at the Crystal-
lographic Institute in Freiburg [Eye81, Eye84], with half-axes 80mm
and 90mm. Furnaces with the same internal geometry have been flown
on several sounding rocket campaigns (TEXUS, module TEM
02-ELLI) and the Spacelab mission D1 (MEDEA-ELLI, old version).
Compare fig. 2-8. Also shown is an ampoule for the FZ growth of Si.
companies, NEC and Tsukuba Asgal in Japan, and the Moscow Power
Engineering Institute in Russia. Practically all mirror furnaces employ
catoptric, not refractive or catadioptric elements because the geometric
efficiency of lenses is quite limited [Eye77] and the heat exposure of the
optical elements excludes most refractive materials except fused quartz
or sapphire.
Either ellipsoidal or parabolic mirrors (or a combination of both) are
used to focus the light from one or several lamps as shown in fig. 2-5b.
Monoellipsoidal mirrors as in figs. 2-7/2-8 have a better efficiency
because nearly all of the light is focused onto the sample; parabolic
mirrors do not use the radiation not reflected by the first mirror shell, but
2. The floating-zone process 17
Fig. 2-8: The monoellipsoid mirror furnace ELLI, half-axes 80/90mm.
Left: Laboratory version plus an additional mirror shell at the bottom.
Right: Module TEM02 for experiments on sounding rockets (TEXUS
program), built by the aerospace company ERNO.
have the advantage that the distance between the foci is variable. A
somewhat better efficiency of parabolic mirrors or of ellipsoidal reflec-
tors using only partial ellipsoids [Ars73a, Ray88] is possible by the
introduction of an additional hemispherical mirror as shown in fig. 2-9.
Another division can be made between furnaces where the sample axis
and the main axis of the furnace coincide, and furnaces where the main
axis is at 90
o
to the sample axis. The former concept as shown in figs.
2-5b, 2-7 and 2-9 allows a very good rotational symmetry of the radia-
tion field (i.e. the thermal field), but the accommodation of different
translation mechanisms for the feed rod and the crystal is quite difficult
(e.g. see the construction in [Bed84]) and the maximum processing
length is usually limited to a value smaller than the distance of the two
foci. In this type of furnace, ampoules are necessary to fix the feed rod
and the growing crystal (fig. 2-7). With the second type of construction,
the processing length is only limited by the translation mechanisms, but
the thermal symmetry is degraded considerably, making crystal rotation
a necessity. Often several mirrors are combined to alleviate the thermal
asymmetry as in the double-ellipsoid mirror furnace in figs. 2-10 and
18 2. The floating-zone process
Fig. 2-9: Image furnace with two
parabolic mirrors and a
hemispherical additional mirror,
1/4 the diameter of the main
mirrors, to increase the solid
angle from W=5.9sr to 11.2sr.
The additional mirror, if actively
cooled, also shields the top of the
sample from direct radiation,
thus improving the axial
temperature gradient.
2. The floating-zone process 19
Fig. 2-10: Double ellipsoid mirror furnace MHF [Eye77, Eye79],
developed at the Crystallographic Institute in Freiburg, with half-axes
80mm and 90mm. A furnace with the same internal geometry has been
flown on the Spacelab missions FSLP (MSDR-MHF) and D1
(WL-MHF). Compare also fig. 2-11.
2-11; this allows also an increase in available heating power. For ellip-
soidal mirrors, however, the good efficiency is reduced because the
radiation directly emitted from one lamp into a different ellipsoid is not
focused and lost. A limitation to two joined ellipsoids has been shown to
be the best compromise [Eye77, Eye79]. A furnace where a part of an
ellipse is not rotated around the major axis but an axis perpendicular to it
(fig. 2-12) can also be regarded as part of this group, or as a cross
between a ring resistance heater and an image furnace. The rotational
symmetry of such a furnace should be quite good, but, probably due to
the manufacturing difficulties of making and supporting the heating
element and the mirrors, only a few constructions have been reported to
date [Dav78, Quo93]. The latter, used for the crystal growth of Ge and
Bi
12
GeO
20
, employed a Super Kanthal resistance heating element,
because it does not need as many mechanical supports as a filament.
Mirrors are usually machined from metal, e.g. aluminum alloys (such
as Al with 3% Mg), brass, or steel. The inner surface has to be polished
and can either be used as is (aluminum alloy mirrors) or is electrolyti-
cally coated with gold (reflectances see fig. 2-13). Silver, although
20 2. The floating-zone process
Fig. 2-11:
Laboratory
version of the
double-ellip- soid
mirror furnace
(MHF-L).
having the best reflectance, is avoided due to its unfavorable tarnishing
properties.
If processing in air is not possible, the processing atmosphere can be
provided by the furnace volume itself in the case of closed mirror
furnaces with vacuumtight feedthroughs, or by additional transparent
containers, e.g. ampoules or fused quartz tubes. Transparent pressurized
vessels for up to 10
7
Pa have been reported [Bal81].
2. The floating-zone process 21
Fig. 2-12: Schematic
view of a float-zone
furnace with ellipti-
cal reflector. Such a
furnace was devel-
oped and built by
CANMET for the
growth of Ge and
Bi
12
GeO
20
crystals
[Quo93]. It employs
a MoSi
2
(Super
Kanthal) heating
element.
Fig. 2-13: Reflectivity of
several mirror materials
from 200 to 4000nm
(after [Nau87]).
Although some of the first furnaces developed employed carbon arcs
[Koo61], the light sources mostly used today are either tungsten halogen
lamps of the order of 0.5-1.5kW (fig. 2-14) or Xenon arc lamps up to
10kW for high power requirements [Kit77, Bal81]. In the latter case, the
spectral distribution of the radiation has to be taken from the manufac-
turer's data. The spectral intensity I of thermal radiators such as tungsten
filaments is given by :
(2-I) I
(
, T
)
=
(
, T
)
$ I
b
(
, T
)
where e=spectral and temperature dependent emissivity of the material,
=frequency, T=absolute temperature, and I
b
=spectral intensity of a
black body radiator given by Planck's law:
22 2. The floating-zone process
Fig. 2-14: Tungsten halogen lamps
typically used in mirror furnaces.
Left: Special development (Sylva-
nia A 708, 36V, 450W). Right:
Commercial type normally used
for studio lighting, available from
several vendors (FEL1000, 120V,
1000W).
(2-II) I
b
(, T) =
2$h
c
2
$

3
e
h$
k$T
1
where h=Planck's constant, c=speed of light, k=Boltzmann's constant,
other symbols as above.
In all cases where lamp bulbs made of fused quartz are employed,
wavelengths shorter than 0.2m or longer than 4m are cut off. For
borosilicate glass the transparent region ranges only from 0.35 to 2.5m.
The filaments or arcs should be as small and isometric as possible, as
the focusing properties degrade rapidly for nonfocal/nonparaxial rays
(fig. 2-15) due to the strong coma in parabolic and elliptic mirrors
[Ray88]. Optical aberrations are more pronounced for strongly curved
surfaces, i.e. for ellipsoids with an axes ratio << 1 [Eye77, Eye79] or
paraboloids with large geometric coefficients (fig. 2-15). Point focus
geometries such as standard ellipsoid or parabolic optics limit the
maximum sample diameter to a few cm. By deviation from the ideal
elliptic shape this can be partially counteracted, introducing a so-called
ringfocus where the energy maximum is not a point, but an annulus
optimized for a given sample diameter [Kra85, Len90, Dan96]. This
concept has been employed in the ellipsoid-based "Automatic Mirror
Furnace" (AMF), which was used on the unmanned space platform
EURECA [Dan96], and in the paraboloidellipsoid combination furnace
MEDEA-ELLI (new version, fig. 2-16) flown on the Spacelab mission
D2 in 1993 [Len90, Cr94b, Dan94, Her95], and the SPACEHAB-4
mission in 1996. The maximum sample size and the maximum tempera-
ture are of course also dependent on the absorption and reflection coeffi-
cients of the solid and liquid material.
2. The floating-zone process 23
Feedback control of image furnaces is uncommon, because no usable
temperature signal can be obtained from the zone without difficulties
[Eye77, Eye81]. Theoretically, pyrometric measurements can be made in
the far infrared where the radiation from the lamp is cut off by the fused
quartz bulbs , but this would exclude any ampoules or fused quartz tubes
around the sample or fused quartz/glass windows in the furnace. A
24 2. The floating-zone process
Fig. 2-15: Focusing properties of differently shaped parabolic and
ellipsoidal mirrors for focal rays and rays at axial positions 5mm
from the focus; shown are three rays 15
o
apart on each side for each
position. f is the distance from the focal point to the apex of the mirror.
The progressive coma for strongly curved geometries is clearly visible.
chopper in front of the lamps is also not practical [Eye77]. The zone is
usually controlled by visual observation of the zone and manual regula-
tion of the power. If automatic processing must be used as in unmanned
space flights, an optimum parameter set (power/translation/rotation) is
first established by test runs and then executed automatically. A control
loop regulates either the lamp power (or voltage if the lamp resistance
can be assumed to be time independent) or the light intensity measured
by photodiodes pointing at the filament. Light intensity control takes into
account changes of the light output not only related to voltage fluctua-
tions and filament resistance, but also to the discoloration of the lamp
2. The floating-zone process 25
Fig. 2-16: Paraboloid -
ellipsoid mirror furnace
developed by the
aerospace company
Dornier (MEDEA- ELLI,
2nd version). The foci of
the two outer parabo-
loids coincide with the
foci of the center ellip-
soid. The lower focus is
an annulus of 20mm .
This furnace was
successfully used on the
Spacelab mission D2
and, under the name
CFZF, on the
SPACEHAB-4 mission.
bulb, or the higher light intensity under microgravity, caused by the
absence of convective gas cooling of the filament [Eye84]. Due to the
very nature of closed mirror furnaces, however, the photodiode signal
can be quite susceptible to changes of the light reflected back from the
sample towards the lamp, such as changes of the zone shape. Light inten-
sity measurement devices employing diffusers and beamsplitters are
possible in open mirror furnaces and are better suited for control [Bal81].
With modern computing techniques, a control loop using the zone shape
or height determined by image analyzers might be possible, but has not
been reported yet. With reflecting samples, the many reflections and
backreflections in a mirror furnace (fig. 2-3 and frontispiece) make
automatic detection of the interfaces difficult.
The power of incandescent lamps can of course be varied between
zero and full power, but most arc lamps allow changes of the light inten-
sity only in a small power range (usually 3/4 to full power). In this case,
or when a constant color temperature is desired with incandescent lamps,
the light flux can be controlled by an aperture in certain geometries (see
[Bal81]).
Temperature gradients are mainly determined by the geometries of the
furnace, the filament, and the sample, as well as the optical and thermal
coefficients of the solid and liquid sample material. The direct radiation
of the lamp onto the sample top in monoellipsoid furnaces and the focus-
ing properties lead to a considerable flattening of the temperature gradi-
ent at the upper interface (fig. 2-17). A small absorber or reflector
mounted between the tips of lamp and sample like the hemispherical
mirror in fig. 2-9 can reduce this effect. It needs to be actively cooled,
though, because otherwise it will heat up and emit radiation itself
[Wat94]. A small (1-5mm) defocusing of the lamp towards the apex of
the ellipsoid also steepens the temperature profile between the focus and
26 2. The floating-zone process
the center of the ellipsoid (fig. 2-17), because it diminishes the amount
of defocused rays coming from parts of the filament located nearer to the
furnace center (green rays in fig. 2-15) in favor of rays coming from near
the apex (blue rays in fig. 2-15). The latter are better focused, but can
lead to a second focus below the original one for certain defocusing
distances [Dol94]. Paraboloidal mirrors with moderate curvature allow a
steeper axial gradient at the expense of efficiency. Trying to get the best
of both concepts, a mirror furnace using a combination of two parabo-
loids and an ellipsoid (fig. 2-16) has been built by the aerospace
company Dornier [Len90]. It was successfully used on the Spacelab
mission D2 in 1993 for the floating zone growth of GaAs crystals (mp:
2. The floating-zone process 27
Fig. 2-17: Numerical simulations of the axial temperature profiles in the
ELLI furnace (fig. 2-8) with the A708 lamp at 100W for a graphite
sample of 15mm diameter and 140mm length. The conductive and
convective heat transport by the surrounding argon is taken into
account, from [Wat94].
1238
o
C) with 20mm diameter at 600W lamp power [Cr94b, Her95], and
for several other materials on the SPACEHAB-4 mission in 1996.
Temperature gradients are also determined by the reflectance, trans-
mittance and emittance values of the materials in relation to the radiation
spectrum on one hand and by thermal conductivities, dimensions,
convection and the latent heat of melting/solidification on the other hand.
A further description is given in section 2.2.2. It should be noted,
however, that the position of the temperature maximum in mirror
furnaces, sometimes called "thermal focus", is usually not located at the
geometric focus, but a few mm towards the center of the ellipsoid.
One important peculiarity of radiation heating, especially image
furnaces and laser heating systems, is the feedback between the heating
power absorbed and the change of absorption and reflection coefficients
28 2. The floating-zone process
[Nas90] 2
0.083-
0.111
0.09 0.07-
0.08
Y
3
Al
5
O
12
/YAG(633nm)
[Nas90] 0.5 0.01-
0.1
0.06 0.04 Al
2
O
3
/Sapphire(633nm)
[CRC81] 0.22 0.14 0.78 0.86 Au(650nm)
[CRC81] 0.37 0.35 0.63 0.65 Fe(650nm)
[Cro82] 0.18 0.65 0.82 0.35 Ge(VIS)
[Cel85] 0.28 0.60 0.72 0.38 Si(VIS)
Ref.

l
[mm
-1
]

s
[mm
-1
]
e
l e
s
r
l
r
s Material
(Wavelength)
Table 2-a: Optical material parameters in the solid (index s) and liquid
(index l ) state for several materials. r: reflectivity, e: emissivity, a:
absorption coefficient. The
s
and
s
values are for smooth surfaces.
Value ranges indicate temperature dependent measurements. Note that
the absorption coefficient is used here, not the absorptance equivalent to
the emittance (=emissivity for opaque materials with smooth surfaces).
VIS: visible range of the spectrum.
upon melting. For some materials, these coefficients change considerably
at the melting point (table 2-a). In figs. 2-3 and 2-4 one can easily see the
substantial increase of the reflectivity of silicon and of the absorption
coefficient (color change) of GGG upon melting, respectively.
The first case, i.e. the increase in reflectivity of an opaque substance
upon melting, leads to the formation of a pattern of crystallographically
oriented droplets and solid material (fig. 2-18) on the surface at melting
temperature [Cel84, Cel85, Jac85]. By this, the system adjusts the
macroscopic average reflectivity in such a way that the melting tempera-
ture is maintained despite the changes in absorption coefficient at the
phase transition. In other words, between the point where the surface
starts melting and the point where the whole surface is molten, a substan-
tial increase in heating power is necessary to allow for the reduction of
the absorptivity, in addition to the latent heat required. The droplets at
the interface are not stable, but move and coalesce in the temperature
2. The floating-zone process 29
Fig. 2-18: Droplet formation on melting silicon rods (8mm ) in a
mirror furnace. Left: free surface, right: covered by SiO
2
, showing
triangular melt areas following crystallographic directions.
gradient due to surface tension effects, and due to gravity. This often
gives the impression, especially with high translation rates at the feed
rod interface, that the material is "boiling", and can introduce some
irregular vibrations of the zone. These effects are enhanced for materials
where superheating of the solid is possible, such as many semiconduc-
tors [Wen78]. Apart from the movements, this change of reflectivity is
advantageous in general in that it leads to a self-stabilization of the
system; it damps the effect of a perturbation or an asymmetry of the
radiative flux, unavoidable in real systems, on the energy flux into the
sample (i.e. on the zone height and shape, the temperature distribution).
Materials with a higher absorption coefficient of the melt than of the
solid such as many oxides show the opposite effect. Upon the formation
of the melt zone, the power must be decreased, and asymmetries in the
30 2. The floating-zone process
Fig. 2-19: Axial incident power distribution at the surface of a GaSb
sample of 10mm in the TEM-02 ELLI furnace at 90W total power,
calculated with and without secondary radiation. From [Wat94].
external radiation/temperature field can be amplified considerably.
Constant attention is necessary for the control of these melt zones.
The complex interplay between the different material parameters,
temperatures, convective flows and the geometry as listed in table 2-b
leads to difficulties in determining the temperature fields in mirror
furnaces. Due to the recent progress in computing, numerical simulations
by finite element methods are able to predict reasonably well some
aspects of the process
b
; with some simplifications (i.e. only one
2. The floating-zone process 31
b
The results in [Kai93, Dol94, Wat94] were obtained by combining a program
(ELLI, see [Dol94]) calculating the radiation field of the mirror furnace (including
secondary radiation, multiple reflections, diffuse reflections, wavelength as a
function of the reflectivity) with the commercial finite element program FIDAP for
a global numerical simulation of the temperature field in the furnace, including the
heat conductivity of the furnace atmosphere.
x
Latent heat
x x x
Convective
heat transport
x x x x x x
Thermal
conductivity
x x x x
Transmittance
x x x x x
Reflectance
x x x x x
Emittance/
Absorptance
x x x x x
Geometry
Lamp
(filament/
bulb/gas
filling)
Furnace
atmosphere
Furnace Ampoule
or
structural
parts
Crystal/
feed rod
Melt
Zone
Table 2-b: Parameters influencing the temperature field in a mirror
heating system; except for the geometry and the latent heat, the parame-
ters themselves may be temperature dependent.
reflection at the mirror) analytical methods are also a possibility [Riv92,
Hay96). For example, numerical simulations have shown that the secon-
dary radiation, i.e. the radiation emitted by the sample, cannot be
neglected in calculating the temperature field in mirror furnaces, as
shown in fig. 2-19 [Kra85, Dol94, Wat94]. A similar result was obtained
by a recent analytical study [Hay96]. Numerical simulations as well as
analytical calculations depend, however, on the availability of reliable
thermophysical data. For a lot of systems, even well-known materials
such as silicon, these are not available with the necessary accuracy
(compare section 3.1.2). Therefore, a check against experimental results
is nearly always necessary.
32 2. The floating-zone process
2.1.1.1c Laser heating
Laser heating shares many aspects with the image furnaces described
above. This includes the good visual control, the in principle unlimited
temperature range, no general limitations for the processing atmosphere,
and the effects associated with the change of reflectivity and absorption
coefficient at the melting point. Due to their monochromatic nature, and
in contrast to mirror furnaces, laser heaters allow straightforward
pyrometric temperature measurements. Secon- dary radiation does not
influence the temperature profile considerably unless radiation shields
are used. Automatic diameter control, e.g. with a second laser at a differ-
ent wavelength [Fej84], is also easier than in mirror furnaces. The energy
efficiency of laser furnaces, however, is not very good compared to
mirror furnaces, starting with a low electrooptical conversion efficiency,
e.g. 10% for a CO
2
laser [Car91]. The available optical power is then
further reduced in the optical system (beam expander, mirrors) by
2. The floating-zone process 33
Fig. 2-20: Axicon
based laser heating
system for pedestal
growth, employing
only catoptric
elements to produce
a ringfocus. A:
axicon mirror, P:
parabolic mirror.
After [Fej84] and
[Fei86].

reflection and absorption losses. Typically, 4-10 optical surfaces are
necessary in advanced systems. Additional pre- and afterheaters are
sometimes used to reduce the necessary laser power. Although all types
of lasers providing the power at an appropriate wavelength might be
used, the cw CO
2
laser with a wavelength of 10.6 m is the predominant
type. This is due to the fact that oxides, the material group where laser
heating furnaces are most often applied, are opaque at this wavelength.
The same reason precludes the use of this laser for standard ampoule or
tube materials (SiO
2
, Al
2
O
3
) as sample containment if processing in an
oxidi- zing atmosphere is not possible. Therefore lasers with wavelength
in the VIS or near IR, such as YAG-Nd
3+
lasers with =1.06m, are also
utilized.
The use of lasers for floating-zone growth was started in 1969 by
Eickhoff and Grs [Eic69] with the growth of ruby crystals and has
continued over the years, see e.g. [Gas70, Tak77, Bur77, Gur78, Kim79,
Dre80, Elw85, Car91, Che95, Che96]. The main application in recent
years has been for pulling optical single crystal fibers by the pedestal
method, e.g. [Fej84, Fei86, Fei88, Tan88, Nas90, Til91, Yan91, Ima95].
Early designs employed a single laser or several lasers directed at the
zone (sometimes with beamsplitter and mirrors, see e.g. [Gas70]), result-
ing in strongly asymmetric temperature profiles. For a good rotational
symmetry, axicon optics as shown in fig. 2-5 or fig. 2-20 are employed
[Fei86, Car91]. A catoptric focusing system as in fig. 2-20 does not pose
any difficulties for cooling the optical elements and, at the CO
2
laser
wavelength, more materials than the ones listed in fig. 2-13 are available
as mirrors (e.g., molybdenum has a reflectivity of 98% at this
wavelength). Another possibility is the use of a catadioptric system as
shown in fig. 2-5. The refractive elements can be made of GaAs, ZnSe
(Irtran4) for =10.6m; Si or Ge are not suitable for high power CO
2
34 2. The floating-zone process
lasers due to absorption bands (Si) or thermal runaway (Ge) [Kar93].
The high refractive index of these materials (n=2.43 for ZnSe, n=3.28 for
GaAs at 10.6m) leads to a considerable loss of power by reflection
(17% for ZnSe) and makes antireflection coatings a necessity. An analy-
sis by Carlberg [Car91] showed that only 50% of the power leaving the
laser is absorbed by the sample (Al
2
O
3
), and 900W of electrical power
(equivalent to 58W laser power reaching the sample) was needed to form
a floating zone in 10mm LiNbO
3
(mp: 1260
o
C) rods. Tilting of optical
elements might be used for moving the zone without sample translation
[Bag86].
The axial temperature gradient in laser heated floating zones is
normally rather high (up to 1000K/cm [Fei88]) if the profile of the laser
beam is maintained or focused by the optical system. This can be advan-
tageous with respect to high possible growth rates [Fei86, Fei88], but
also introduces higher thermal stress in the crystal and strong, time-
dependent thermocapillary convection in the zone (chapter 3). If these
disadvantages outweigh the benefits of a high pulling rate, the axial
gradient (as well as the interface curvature) can be changed by additional
pre- and afterheaters, thermal shields around the sample, defocusing of
the beam profile by the optical system, the use of several laser/axicon
systems for producing concentric ring beams [Car91], or a combination
of these approaches.
2. The floating-zone process 35
2.1.1.2 High frequency heating
Heating by high frequency (HF) is one of the major methods of
floating-zone growth. HF heating with frequencies between several
100kHz and a few 100Mhz is referred to as radio frequency (RF)
heating, with frequencies from 0.5 to several GHz as microwave heating.
It is based either on induction heating for conductive or ferromagnetic
materials or on capacitive heating for dielectrics such as organic materi-
als, the former being the predominant type of operation. Industrial float-
zone processing is almost exclusively done by RF induction heating.
In induction heating, the charge is surrounded by the HF carrying coil
and the heating effect is due to the resistance heating by the induced
eddy currents in the material. The setup is similar to an electric trans-
former with the sample as a single turn secondary coil. Additional
heating due to magnetic losses (hysteresis losses) occurs only in ferro-
magnetic materials below the Curie point and is usually of no practical
importance for crystal growth. Typical frequencies range from 400kHz
to 5MHz and the maximum resistivity of the sample should be below
100cm. For materials such as undoped semiconductors or salts, where
the conductivity of the material is sufficiently high only above certain
temperatures, the initial heating must be done by an additional heater,
e.g. by radiation heating. This can also be achieved by introducing
temporarily an additional charge (e.g. a metal or graphite ring) which in
turn heats the sample by radiation and/or heat conduction until the
sample conductivity is high enough to absorb a sufficient amount of the
induced energy.
In principle, processing can be done both under vacuum or with a gas
atmosphere. When using a gaseous atmosphere, e.g. to reduce the evapo-
ration of material, the voltage between the coil turns and/or the coil and
2. The floating-zone process 35
the - usually grounded - sample should not exceed the breakdown
voltage for the given gas species and pressure to avoid corona discharges
or arcing. Argon and hydrogen, in contrast to helium, have suitably high
breakdown voltages; in the case of Ar, a small addition of N
2
is helpful
in this respect [Boh94]. At the cost of efficiency, the inductor coil and
the sample can be separated by a nonconducting container, e.g. to avoid
condensation of evaporated material onto the cooled coil.
Several other boundary conditions must be considered in the heater
construction and choice of frequencies, namely zone dimensions, heating
efficiency, skin depth, and the electrodynamic pressure.
Heating efficiency is determined by the coupling between the sample
and the primary coil; the coil shape should preferably follow the sample
shape as closely as possible. Furthermore, the resonant resistance of the
oscillator should be matched to the impedance of the sample, the former
given by (after [Sha80a]):
(2-III) R =
530$Q
C$
0
where Q=circuit quality factor=reactance/resistance, C=circuit capaci-
tance and
0
=resonance frequency= , L being the circuit 1/
(
2 $ L $ C
)
inductance.
Due to the self induction effect, the current distribution is not constant
across the sample diameter, but follows
(2-IV) I = I
0
$ e
x
sk
36 2. The floating-zone process
where I
0
=current at the sample surface, x=depth below the sample
surface,
sk
=skin depth, the depth at which the current is 1/e of the
surface current.
The skin depth is dependent on the frequency as well as the resistivity
and permeability of the sample and can be calculated by
(2-V)
sk
=
!
$$
0
$
r
where =resistivity, =frequency,
r
=relative permeability (close to 1
for most melts),
0
=absolute permeability=4 10
-7
Vs/Am.
For high frequencies (MHz), the heat is thus generated in a very thin
layer and then distributed by heat conduction and convection. For silicon
and a frequency of 2.4MHz,
sk
is about 292m [Boh94]. Consequently,
the sample resistance is no longer inversely proportional to the cross
section, but to the circumference of the sample. For the same reason,
tubes instead of wires can be used as primary coils without increasing
resistance losses; in addition, they allow very efficient water cooling.
Arcing problems and sufficient cooling of the primary coil must also
be considered in the choice of the operating frequencies [Gup78],
because for a given power the RF voltage goes up with
3/4
and the
current goes down with
1/4
[Kel81]. Fig. 2-21 shows some possibilities
for coil geometries: a) is the basic single turn coil, b) to f) are different
possibilities for multiple turn coils, which allow higher currents in the
sample in proportion to their number of turns. Case 2-21b, sometimes
called a pancake coil, allows a better concentration of energy at the
expense of efficiency, whereas 2-21c gives better coupling, but the
2. The floating-zone process 37
induced current is spread over a larger area. Case 2-21f is similar to
2-21c, but makes use of an additional watercooled concentrator to focus
the induced currents.
An important aspect of induction heating is the presence of the
Lorentz force (eq. 3-XXXV), which for the case of induction heating can
be written as [Mh83]:
(2-VI) F
L
=

2
H
2
+
(
H
)
H
with H=magnetic field strength, =
0

r
=permeability.
The resulting electrodynamic pressure, a repulsive force between the
induction coil and the zone, can be used to support larger zones than
usually possible. Fig. 2-21d shows a possible shape for a levitating coil.
The electrodynamic pressure is inversely proportional to the square root
of the frequency (eq. 2-XXXVI, section 2.2.3), so lower frequencies
would allow a larger levitating effect. If, due to other considerations, the
possible frequency range for the heating coil is limited to higher values
[Gup78, Kel81], a setup with two independent coils, one mainly for
heating, one mainly for levitation, can be used (fig. 2-21e).
The electrodynamic pressure is also utilized in a special coil configu-
ration called needle-eye technique (fig. 2-22). Industrial FZ growth of
silicon is almost exclusively done with this setup. The large single turn
"pancake" coil has an inner diameter which is considerably smaller than
the diameters of both the feed rod and the growing crystal. It usually
consists of two parts, electrically separated by a ground connection (gray
section in fig. 2-22), to halve the voltage between the coil and ground,
and thus reduce arcing problems [Kel81]. For the same reason, compara-
tively high currents (above 1000A [Kel81]) are employed, necessitating
very good cooling. For silicon, the electrodynamic pressure of this
38 2. The floating-zone process
arrangement allows an approximate doubling of the length of a floating
zone to values of about 30mm. In addition, the diameter of the liquid is
considerably reduced at the center by being compressed through the
"needle-eye", thus enabling absorption of energy not only at the circum-
ference of the rods as it would be the case in the arrangements of fig.
2-21, but also close to the solid-melt interfaces. The crystal interface
shape is mostly concave or w-shaped with this arrangement and the
aspect ratio external zone height/zone diameter can be much smaller than
usual (compare chapters 2.2.2 and 2.2.3). For FZ-silicon, the maximum
industrially produced diameter is now 150mm (6") [Zuh89].
The coil inner diameter and the coil contours influence the interface
shape (and the radial segregation) to a large extent [Kel81, Mh83,
Rie95, Mh95]. For disk-shaped coils with rectangular cross section,
2. The floating-zone process 39
Fig. 2-21: Different coil configurations for RF heating. a: single-turn
coil b: pancake coil c: multiple turn coil d: levitating coil e: two
separate coils, the upper single-turn coil for heating, the lower three-
turn coil for levitation f: coil with concentrator. After [Sch64, Jon74,
Sha80a, Sha80b]; compare text.
thinner disks appear to be advantageous [Kel81]. Steps at the coil
bottom, as well as coils with a wedge-shaped cross section (fig. 2-22)
tend to flatten the lower interface due to an afterheater effect [Kel81].
Further optimization leads to coil shapes with rather complicated cross
sections [Rie95]. Another possibility to influence the interface shape -
apart from the introduction of separate afterheaters - is a deviation from
circular symmetry, by moving the crystal axis laterally with respect to
the feed rod and zone axis (eccentric needle-eye float-zone technique
[Kel81]), by moving the coil to an eccentric position with respect to the
crystal and feed rod [Sch89], by using elliptic shapes for the coil
openings [Sch89], or a combination of these.
In addition to the control concepts mentioned in section 2.1, one can
also use the feedback signals from the tuned heater circuit in the case of
induction heating, because any changes in the shape and size of the
40 2. The floating-zone process
Fig. 2-22: Schematic view of a single turn watercooled RF heating
coil for needle-eye floating-zone growth.
crystal as well as the zone change the inductance of the system and thus
the position of the resonance curve (fig. 2-23). For a given system,
thermal or mechanical self-stabilization can be achieved by an appropri-
ate choice of the frequency [Kel81]: For materials with a negative
temperature coefficient of the resistivity, a fixed workpoint at
a
on the
inductive slope of the resonance curve leads to thermal stabilization,
because a temperature increase leads to a higher conductivity and thus to
a smaller inductance, which shifts the resonance frequency
0
to higher
values (black curve -> gray curve in fig. 2-23). By this shift, the coil
voltage (i.e. the power) goes down, moving from a to a'. For materials
with positive temperature coefficients of the resistivity, a workpoint on
the capacitive slope gives thermal stabilization. In both cases, a mechani-
cal stabilization is possible with a fixed workpoint at
b
on the capacitive
slope; any movement of the melt or crystal towards the coil increases the
power (the voltage goes from b to b') and thus the electrodynamic
pressure at this point .
Capacitive heating of insulators by high frequency electromagnetic
radiation utilizes the interaction of the alternating electric field with the
dielectric polarization of the material. Heating is only possible when, due
to relaxation of dipoles, a phase difference between the electric field
vector and the polarization vector is present. The amount of heating is
governed by the loss factor

of the dielectric, which is the imaginary


part of the frequency dependent complex dielectric constant :
(2-VII)
&
= i $

Possible contributions to the polarization of a material are electronic


polarization, atomic polarization, dielectric polarization and, for hetero-
geneous systems, the so-called Maxwell-Wagner polarization due to
2. The floating-zone process 41
charge build-up at interfaces. The contribution of the first two to the loss
factor are in the VIS and IR; this leaves the dielectric polarization losses
(sometimes called reorientation losses), and conductivity losses for the
heating of a single phase system with high frequency. The useful
frequency range is considerably higher than that of induction heating,
ranging from radio frequencies of 30MHz [Sha80a] to microwaves of
several GHz [Sch64, Met88]. The heating efficiency is given by the total
power loss w in the dielectric, which is given by [Sha80a]:
(2-VIII) w = U
2
$ C$ *$

42 2. The floating-zone process


Fig. 2-23: Schematic voltage (power) - frequency diagram for two
different resonance curves of a tuned RF heater circuit (after [Kel81]).
Workpoint a gives thermal stabilization, workpoint b mechanical stabili-
zation of the zone. See text for details.
where U=voltage, C=capacitance, =rotational frequency,

=dielectric
loss factor.
For a given frequency, the maximum power for capacitive heating, i.e.
the maximum voltage, is given by the electric breakdown limit of the
material. For frequencies in the microwave range, i.e. above 0.5GHz, a
wired circuit such as a coil cannot be used. The microwave radiation,
usually generated by a klystron or magnetron, is transferred to the
sample by a waveguide. A concentration of the microwave energy is
possible by designing the growth chamber as a resonant cavity [Met88].

It should be mentioned that in magnetic materials a magnetic loss
factor

of the complex permeability * might contribute to heating
[Met88]. The effect, not to be mistaken with hysteresis losses, is caused
by domain wall and/or electron spin resonance in the RF and microwave
range of the spectrum.
2. The floating-zone process 43
2.1.1.3 Electron beam and plasma heating
Both methods use mainly charged particles, i.e. electrons and/or ions
to heat the sample; their principles are illustrated in fig. 2-24. The main
practical difference is that electron beam heating requires vacuum,
whereas an ambient gas is necessary for all plasma heating methods.
2.1.1.3a Electron beam heating
In electron beam heating, the sample is heated by the absorption of
kinetic energy of electrons emitted from a cathode and then accelerated
by the applied voltage to the sample anode (fig. 2-24). Since its intro-
duction in 1956/1957 by Davis and Calverley [Cal57], the main use in
float-zone processing in the last 40 years has been the crystal growth
(and zone refining) of refractory metals and alloys with high melting
points, see e.g. [Neu62, Sel64, Mau68, Hay78, Jur82, Gle89, Jur90a,
Jur90b, Sem95, Liu96]. The method is often called EBZM (electron
beam zone melting) and requires operation under a vacuum better than
10
-4
mbar. For this reason, materials with a considerable vapor pressure at
the melting point are not very suitable, though an additional purification
effect is often achieved by the outgassing of volatile impurities during
processing [Sch64]. In the usual setup, with the sample as anode, the
sample material must be conducting, excluding insulators and most
(undoped) semiconductors. This disadvantage can be overcome by the
use of an additional grid as anode around the sample.
Typical voltages are several kV, the upper limit set by the generation
of X-rays, with currents of the order of 10
-2
-10
-1
A. In most cases, the
cathode is on ground potential and the positive voltage is applied to the
44 2. The floating-zone process
sample [Boh94]. Electron beam heating can be highly efficient, with
over 99% of the cathode current arriving at the sample [Sha81a].
Current-voltage characteristics are similar to those of a vacuum diode:
Up to the saturation current, the current is space charge limited, with the
current density j given by
(2-IX) j l
0
$
2$e
m
$
U
3
2
x
2
where
0
=permittivity of vacuum, e=electron charge, m=electron mass,
U=voltage, x=distance between the electrodes.
The saturation current is given by the Richardson equation:
(2-X) j
sc
= A$ K$ T
2
$ e

W
e
k$T
2. The floating-zone process 45
Fig. 2-24: General representation of the principles of electron beam
heating (a) and plasma heating (b) in FZ growth. In electron beam
heating, vacuum is necessary and the zone as anode is heated by the
accelerated electrons. In plasma heating, requiring an ambient gas,
there is no macroscopic charge and the heating is by excited electrons
and ions from the plasma. See text for details.
where A=area, K=material constant ( 60 Acm
-2
for metals), T=filament
temperature, W
e
=emission work function (4.54eV for tungsten),
k=Boltzmann's constant.
The most simple arrangement uses a ringwire as cathode surrounding
the zone (fig. 2-24), often made from tungsten or the same material as
the crystal to avoid contamination. If the latter is not possible, contami-
nation by filament material can be a problem with this setup at higher
temperatures. Similar to vacuum triodes/pentodes or electron micro-
scopes, more sophisticated electron guns (see e.g. fig 2-25) use modula-
tor grids, additional focusing and accessory electrodes, as well as
electron lenses to reduce this problem and to allow better control of the
electron beam. The cross section of the electron beam can be reduced to
an area of m
2
to produce very high power densities (up to 10
5
kWcm
-2
[Sha81a]) and high temperature gradients. Control of the power, i.e. of
the emission current is achieved either by controlling the cathode-anode
voltage or the filament temperature (maximum working temperature is
0.9 mp). Control of the emission current can be complicated by a
positive feedback due to the outgassing of material from the sample or
the cathode: The additional ions increase the current, which in turn
increases the sample temperature leading to even stronger outgassing.
This cycle can end in a voltage breakdown by glow or arc discharge.
Temperature measurements by pyrometry are possible and have for
instance been used to determine the temperature gradients in an electron-
beam heated floating zone of refractory metals [Jur82, Jur90a, Jur90b].
2.1.1.3b Plasma heating
Setups requiring or permitting operation in a gas atmosphere allow the
use of a plasma as heat source. In a plasma the atoms or molecules of the
46 2. The floating-zone process
gas are excited and ionized to a degree allowing good conductivity, but
there is no macroscopic charge. The heating effect is due to the transfer
of kinetic energy of the electrons, ions and sometimes excited neutral
molecules of the plasma to the sample. A plasma can be characterized by
a) the degree of ionization, b) the operating pressure, c) the plasma
temperature, and d) the method of ionization, e.g. glow discharge,
electric arcs, RF and microwave excitation, electron-cyclotron resonance
(ECR), focused laser radiation and flame heating.
A glow discharge plasma is a low pressure (typically 0.1-1mbar), low
temperature and low degree of ionization (a few %) plasma. The depend-
ence of the current density j on the gas pressure p is given by
(2-XI) j i p
2
The plasma is generated by a suitable voltage between two electrodes.
Often, but not always, one of the electrodes is the sample. Usually the
glow discharge in the chamber is concentrated in two regions, the
cathode fall region with a high electron density due to ion bombardment
2. The floating-zone process 47
Fig. 2-25: Setup of
an electron gun with
additional electrodes
to focus the electron
beam. The contami-
nation of the sample
by tungsten evapo-
rated from the
cathode wire is
greatly reduced by
the larger distance
plus the shielding of
the upper accessory
electrode. After
[Sem95].
of the cathode, and the anode fall region with thermally excited ions and
electrons. Both regions have been used for plasma heating in FZ crystal
growth, see e.g. [Tro62, Cla67, Cla68, Sto70, Bro71]. In addition to this
naturally occurring concentration of the plasma due to space charge
effects, hollow cathodes or anodes allow further focusing of the energy.
According to [Sha81a], typical parameters for a hollow cathode appara-
tus are 1.5-5kV and 0.1-1.5A in the pressure region specified above.
Operational limits of the process are given by the transition to an arc
discharge and sample contamination by sputtering of electrode material
at elevated temperatures.
The electric arc discharge is distinguished from the glow discharge by
the different mechanism of electron emission from the cathode. In the
case of a glow discharge, electrons are emitted mainly due to ion
bombardment, whereas in an arc discharge the higher electrode tempera-
ture leads to thermal emission of electrons. Thus the necessary voltage
for the discharge is reduced and the current density goes up considerably
(j 0.1Acm
-2
[Sha81a]). In this case, the pressure dependence of the
current density is
(2-XII) j i p
4
3
By using an electric arc, very high temperatures can be attained, but it
is quite difficult to achieve sufficient temperature control. Another
problem is the high degree of contamination by electrode material,
although this can be reduced or overcome by watercooling of the
electrodes [Ger63], or by using electrodes made from the same material
as the crystal [Ver76, Mac88].
A plasma can be generated by RF heating with frequencies above
4MHz at low vapor pressures [Sha81], similar in construction to a RF
48 2. The floating-zone process
plasma torch used for welding. Focusing of the plasma flame can be
achieved by magnetic and electric fields. To initialize the ionization,
temporary additional heating of the gas is necessary, e.g. by inserting a
conducting rod into the RF field. The method is capable of achieving
very high temperatures in excess of 4200K and there are no problems
from contamination; carbides of tantalum, hafnium, niobium and
vanadium have been grown by this process [Sav78, Kum81a, Kum81b].
Similar to excitation by radiofrequency, a plasma can also be generated
by microwaves in low pressure gas atmospheres. The concentration and
positioning of the microwave plasma is possible by a suitable design of
the waveguide and the cavity.
Float zoning with a gas burner, similar to the heating employed in the
Verneuil method, has been scarcely used. Possible combustible gases are
H
2
, propane and butane with O
2
as oxidizer. It is of course only suitable
for processes without necessity of good temperature control and where
contamination by the burning gas or by air is not important. A few
oxides have been grown by this method [Bro64, Tie72].
2. The floating-zone process 49
2.1.1.4 Direct heating
The term direct heating is used for two entirely different processes.
The first one, also termed Joule heating, uses the Joule heat of a current
through the sample as heat source (fig. 2-26). The voltage is directly
applied between the upper and lower ends of the sample. A resistance
gradient is then generated either by controlled active cooling of the
sample ends, by active additional heating of a small part of the sample,
or passively by using a small heat reflector around it [Pfa66]. More
power is dissipated in the hotter part than in the colder part, thus forming
a zone by positive feedback. Joule heating is more often used as
additional heating of the whole sample to reduce the required power of
the main power source, or to reduce thermal gradients and influence the
interface shape [Dor64].
In the second process, a heater is immersed into the melt, but has no
contact with the crystal (fig 2-27). The heat is directly transferred from
the heating element to the melt by conduction. This arrangement is of
course not completely contamination-free and the chemical inertness of
the heater is of utmost importance. It might be argued that for this reason
it is not really a floating-zone process, but it retains the advantage that no
external stress is imposed onto the crystal by an ampoule or crucible.
The first heater of this type was the so-called strip heater (fig. 2-27a),
introduced in 1965 by Gasson for the crystal growth of Nd-doped schee-
lite crystals [Gas65]. A strip (resistance) heater consists of a small
iridium or platinum sheet inserted into the melt zone. Holes in the metal
strip allow the necessary material transport through the heater. This type
of heater has also been used to grow calcite [Bri71, Bel72, Bel76],
Ba
0.65
Sr
0.35
TiO
3
[Hen74], BaTiO
3
[Tur82], and Li
2
CO
3
[Pal82]. In
addition to the strip heater, several other direct heater configurations
50 2. The floating-zone process
have recently been investigated by Lan and Kou [Lan91a, Lan91b,
Lan91c], namely a step heater (fig. 2-27b) and ring heaters (fig. 2-27c,
d). NaNO
3
was used as a model substance for these investigations, with
heaters made from graphite and/or aluminum. The heaters were designed
to act directly as a shaping die, similar to the EFG process. The main
goals were improving diameter control over the standard strip heater
design by reducing the melt creep caused by wetting effects, and a
significant reduction of thermocapillary convection by reducing the free
surface of the zone (see 3.4.3). The step heater (fig. 2-27b) and the
completely immersed ring heater (fig. 2-27c) gave the best shape control
[Lan91a, Lan91c]. The interface shape is of course also influenced by
the heater design and position. For the ring heaters, larger stable zones
than usually possible (see chapters 2.2.1 and 2.2.3) have been reported
and were attributed to the
additional supporting effect of the
ring [Lan91a].
2. The floating-zone process 51
Fig. 2-26: Principle setup of Joule
heating.
52 2. The floating-zone process
Fig. 2-27: Immersed heater systems: a) strip heater, b) step heater, c)
and d) ring heater. After [Gas65, Lan91a, Lan91b, Lan91c].
2.2 Floating-zone stability and control
The success of a floating-zone growth experiment depends to a large
extent on the ability to control the zone size and interface curvatures and
to understand their dependence on the growth conditions. An essential
parameter is the hydrostatic stability of the zone (sections 2.2.1, 2.2.1.1),
where exceeding the limits results in immediate termination of the
experiment. In addition to this static stability, the influence of transient
conditions during the crystal growth process must be considered; this
includes the influence of the growth angle and of volume changes on the
crystal size during the process (section 2.2.1.2). The different variables
determining the zone geometry, such as heater profiles and convective
flows, are discussed in section 2.2.2; additional possibilities for control
through external fields are given in section 2.2.3.
2.2.1 Zone and crystal shape
A floating melt or solution zone is held only by surface tension and
adhesion between feed rod and growing crystal. Under the approxima-
tion that the surface tension is not dependent on position
a
and neglecting
influences by flows in the melt zone and the surrounding medium, the
shape of a floating zone, as that of any liquid volume with free bounda-
ries, can be described by the Laplace (or Young-Laplace) equation:
2. The floating-zone process 55
a
This is of course not true for a real floating zone, but the change in the surface
tension due to temperature or concentration gradients is usually a few % at maximum
and thus introduces only a small error in the static calculation of the zone shape. As
the driving force of thermocapillary convection it has, however, a considerable
impact (compare section 3.1).
(2-XIII) Ap = , $
1
R
1
+
1
R
2
where p=pressure difference at the surface, =surface (interface)
tension, R
1
and R
2
= principal radii (see fig. 2-28).
To put this equation to use one has to take into account the given
geometry, which defines the principal radii (fig. 2-28) by functions and
appropriate surface parameters, as well as any pressures in addition to
the always present capillary pressure, e.g. the hydrostatic pressure under
gravity. For most geometries, no analytical solutions are possible and eq.
2-XIII must be solved numerically.
2.2.1.1 Static stability
56 2. The floating-zone process
Fig. 2-28: Principal radii R
1
(green arrow) and R
2
(red
arrow) from the Laplace
equation (2-XIII) for the floating
zone case. Note that in contrast
to R
2
, the origin of R
1
is not
necessarily located on the
z-axis.

For axisymmetric zones, and taking the hydrostatic pressure due to
gravity into account, eq. 2-XIII can be written as [Cor77a, Cor77b,
Riv79, Teg95]:
(2-XIV)
Ap
0
,

Ap
l
$g
,
$ z =
1
r$
(
1+r
2
)
1
2

(
1+r
2
)
3
2
where z=cylindrical axial coordinate, r=r(z): zone radius at axial coordi-
nate z with r' and r'' as the first and second derivative, respectively,
=surface tension, p
0
=capillary pressure difference between melt and
surrounding gas or liquid at z=0,
l
=density difference between melt
and surrounding gas or liquid, g=gravitational acceleration.
By introducing dimensionless variables, the zone shape can be
described by four parameters [Cor77a, Cor77b], viz:
(2-XV)
r
f
r
c
,
L
r
f
,
V
v$r
f
2
$L
,
p
l
$g$r
f
2
,
= Bo
s
where r
f
=feed rod radius at interface, r
c
=crystal radius at interface,
L=zone length, V=zone volume, Bo
s
=(static) Bond number.
The static Bond number expresses the balance between the destabiliz-
ing effect of gravity versus the stabilizing effect of surface tension. For g
0, as in microgravity conditions, it is close to zero and hydrostatic
pressure can be neglected. Zone (iso)rotation introduces a fifth parameter
similar to the static Bond number to account for the centrifugal force
[Cor77a, Cor77b]:
2. The floating-zone process 57
(2-XVI)
p$c
2
$r
f
3
,
where =angular velocity, other symbols as above.
Using numerical methods to solve the above equations, one can obtain
the shape for a given set of parameters r
f
- r
c
- V - -
l
, e.g. for paramet-
ric studies. This has been done by a number of authors, e.g. [Cor77a,
Cor77b, Riv79, Bou85, Lan90, Mar95, Tat94, Teg95], for a variety of
configurations.
An important special case of float-zone processing is the cylindrical
floating zone (fig. 2-29 top left), i.e. r
f
/r
c
=1, V/(r
2
L)=1 and Bo
s
0
(either g0 ms
-2
as in microgravity conditions, or the zone, i.e. r
l
, is
very small) where the maximum stable zone lengths L
max
is given by:
(2-XVII) L
max
= 2 $ v $ r
This is the famous Rayleigh limit, first observed by Plateau [Pla63]
and theoretically derived by Lord Rayleigh in 1879 [Ray79]. It states that
the length of a cylindrical zone cannot exceed its circumference, because
any sinusoidal perturbation with a wavelength larger than 2 r will
result in a smaller surface and thus a smaller energy than the cylindrical
shape. This relation governs the floating zone process under micrograv-
ity conditions. It should be noted, however, that longer zones are possi-
ble with V>r
2
L, i.e. barrel-shaped zones, and that a zone can break
below the Rayleigh limit if V<r
2
L. The latter is the case if one starts
with a cylindrical sample and the density of the melt is higher than that
of the crystal. In addition to the well-known case of H
2
O, most semicon-
ductors also show this anomaly (table 2-c). It should be noted that for
58 2. The floating-zone process
Bo
s
0 the zone stability is independent of the surface tension value or
any other material parameters.
Under gravity, the resulting hydrostatic pressure changes the meniscus
to a "bottle" shape (fig. 2-29, right column, and fig. 2-30). Therefore the
meniscus angle
M
at the lower interface is increased and the meniscus at
2. The floating-zone process 59
Fig. 2-29: Silicon float-
ing zones under micro-
gravity (left column),
and 1g conditions (right
column) in a monoellip-
soid mirror furnace (see
figs. 2-7 and 2-8). The
g experiment, done on
a sounding rocket
(TEXUS 29), and the 1g
experiment were made
in the same furnace and
with similar parameter
settings (less power was
used for the g experi-
ment, see section 2.2.2).
Shown are the first melt
zone (top row) formed
from a cylindrical rod
of 8mm , and the zone
and grown crystal after
about 5mm of growth
(bottom row); i marks
the melt-crystal inter-
face. The translation
rate was 5mm/min.
the upper interface usually bulges inward (for
r
c
=r
f
). In this case, the limit for a stable zone
length is given by
M
= 90
o
(
if
+
W
), where
W
is the wetting angle of the melt with its own
solid (e.g.,
W
is 33
o
for Si, 30
o
for Ge [Sat80]).
Normally, the maximum possible zone length
under this conditions is considerably smaller
than in the microgravity case.
Assuming flat interfaces and large equal radii
of crystal and feed rod (approximately r
f
= r
c

10mm, depending on ) under gravity, the


above condition leads to the well known
equation [Hey56, Cor77a, Cor77b, Lan90,
Teg95]:
(2-XVIII) L
max
= K
(

M
)
$
,
p
l
$g
where =surface tension,
l
=density, g=gravita-
tional acceleration; K(
M
) is a factor dependent
on the meniscus angle
M
(fig. 2-30).
The dependence of K on
M
with
M
in degrees can be expressed by
[Teg95]:
K
(

M
)
= 2.672 + 1.815 $ 10
2
$
M
7.642 $ 10
5
$
M
2
(2-XIX)
60 2. The floating-zone process
Fig. 2-30: Basic
configuration for a
floating zone under
gra- vity.
M
and
M
are the lower and
upper meniscus
angles,
ic
and
if
the
angles of the crystal
and feed rod inter-
faces at the solid-
liquid-gas trijunction,
resp.
For constant diameter growth, the value of the meniscus angle
M
must
be identical to that of the growth angle
G

b
(tables 2-d, 2-e).
Eq. 2-XVIII was originally found by Heywang [Hey56] in an approxi-
mate calculation; it is often called the Heywang limit. He only consid-
ered the case
M
=0
o
and derived a value of K=2.84. The calculations of
Coriell and Cordes [Cor77a, Cor77b] resulted in K=2.67 for
M
= 0
o
; this
value is corroborated by eq. 2-XIX from [Teg95]. Interestingly, using a
value of
M
=11
o
(the growth angle of silicon) in eq. 2-XIX, results in K=
2.86, close to Heywang's value. Table 2-d lists the surface tension,
2. The floating-zone process 61
b
Note that the growth angle
G
itself is not really a material constant, but depends
on several other conditions, among them the crystal orientation and the interface
curvature (see the following section, 2.2.1.2).
+4.7 [Tay85] 5.87 [Str77] 6.16
rt
CdTe
-12 [Gla69] 6.47 [Ml84] 5.78
rt
InSb
-3.2 [Gla69] 5.85 [Ml84] 5.67
rt
InAs
-5.9 [Gla77] 5.07 [Ml84] 4.79
rt
InP
-7.4 [Gla69] 6.03 [Ml84] 5.61
rt
GaSb
-7.6 [Gla69] 5.71 [Ml84] 5.31
rt
GaAs
-4.8 [Gla69] 5.51 [Mh84] 5.26
mp
Ge
-11.7 [Sas95] 2.57
-10 [Rhi95] 2.53
-9.6 [Gla69] 2.52
[Mh84] 2.30
mp
Si
p
s
p
l
ps
[
%
]
Ref.
l
[gcm
-3
] Ref.
s
[gcm
-3
] Material
Table 2-c: Densities in the solid (
s
) and liquid (
l
, at mp) state of some
semiconductors and the resulting volume change upon melting (rt: at
room temperature, mp: at melting point). All elementary semiconductors
of the diamond lattice type (space group Fd3m) and most of the binary
semiconductors of the sphalerite or wurtzite lattice type (space groups
F$3m and P6
3
mc, respectively) have a negative volume change.
density, growth angle and the resulting K and L
max
according to eqs.
2-XVIII and 2-XIX for several materials. It is evident that except for
silicon, having the favorable combination of a high surface tension and a
low density, L
max
is limited to a few mm under gravity for most materials.
Large growth angles like those of InSb or GaSb are not only favorable
because of their higher stability in transient growth situations (see
section 2.2.1.2); they also increase static stability due to the slightly
higher value of K(
M
). L
max
and the crystal diameter are independent of
each other according to eq. 2-XVIII, which would in theory allow the
growth of large diameter crystals. In practice, this would only be possi-
ble for completely planar or concave interfaces. For thermal reasons
(section 2.2.2), such conditions are usually not possible to attain for large
62 2. The floating-zone process
12.5 2.96 174[Dre80] 3.80[Sat80] 670[Sat80] Al
2
O
3
6.6 290[Nak92]
8.1 3.08 251[Sat80] 6.47[Gla69] 434[Har93] InSb
8.7 3.16 30.72[Teg96] 6.03[Gla69] 450[Teg95] GaSb
7.8 2.93 15 [Hur63] 401[Rup91]
9.9 2.96 17.31.6[Teg95] 5.71[Gla69] 631[Wan90] GaAs
9.3 2.80 73[Wen78]
9.7 2.90 13[Sur76a] 5.49[Gla69] 600[Kec53] Ge
15.1 2.80 6.1-8.1[Teg95] 718[Sas95]
17.1 2.86 11[Sur75] 2.52[Gla69] 885[Har84] Si
L
max
[mm K
G
[
o
]
l
[gcm
-3
] [10
-3
Nm
-1
] Material
Table 2-d: Surface tension , density
l
, growth angle
G
, factor K(
G
)
calculated by eq. 2-XIX, and maximum zone length L
max
for large zones
under 1g, calculated by eq. 2-XVIII for different materials. For other
densities of liquid Si see table 2-c. If several material parameter values
were available, both the best and the worst case for L
max
are shown.
crystals. A good rule of thumb for a lot of growth systems is that the
maximum diameter cannot exceed the zone length without the danger of
forming a solid bridge between crystal and feed rod, i.e. the aspect ratio
L/2r should be equal to or larger than 1.
A concave lower interface, often encountered in needle-eye RF
heating, or through the combination of partially transparent materials
(oxides) and radiation heating, actually stabilizes the zone and increases
the maximum zone length by supporting the lower end of the meniscus
[Kel81]. A concave upper interface, however, increases the hydrostatic
pressure at the lower interface due to the larger liquid column height in
the center and hence acts as a destabilizing influence.
For the cases not covered by eq. 2-XVII and eq. 2-XVIII, e.g. with
intermediate values of r or for zones with r
c
r
f
, L
max
must be calculated
individually. An interesting result of an analysis of meniscus shapes is
the fact that for certain combinations of r,
M
and L several solutions
(fig. 2-31) of the Laplace equation - with different volumes- exist
[Cor77a, Cor77b, Teg95]. Different meniscus shapes also result if the
crystal pulling direction for a given floating zone process is reversed
with respect to the gravity vector [Wil88]. Depending on the particular
system, stable growth is only possible for a certain range of the ratio
crystal diameter/feed rod diameter; this range is actually larger for
pulling upward than for pulling downward [Tat94]. For the pulling direc-
tion being opposed to the gravity vector, the requirement
M
=
G
0
o
,
however, can only be fulfilled for small zones [Dur86], or a substantially
smaller diameter of the growing crystal in comparison to the feed rod as
in the standing pedestal technique. Pulling downward is therefore the
most common case. Stability diagrams for various floating zones (or,
more generally, liquid bridges) can be found in [Cor77a, Cor77b, Riv79,
Lan90, Tat94, Mar95, Teg95]; tables 2 - 4 in [Mar95] give a
2. The floating-zone process 63
comprehensive list of analyses done on the hydrostatics, stability limits,
and dynamics of liquid columns in general. An example of L
max
as a
function of the crystal radius is shown for silicon in fig. 2-32. As a rever-
sal of the usual calculation, one can use the zone shape under gravity to
calculate approximate values of the surface tension for new material
systems [Nak90, Teg95, Teg95a]. Strongly bulging zones increase the
accuracy of this method of surface tension determination, with typical
errors of the order of 5-10%.
In real crystal growth situations, the zone length should be kept well
below the maximum value, because the static stability limit is derived
from the reaction of the surface energy to an infinitesimal small distur-
bance. Any disturbance of finite value, e.g. due to shocks or vibrations,
might easily cause a zone to rupture if its length is close to L
max
. This is
especially true for accelerations perpendicular to the zone axis. Floating
zones are very sensitive to vibrations; GaAs-FZ experiments on the
64 2. The floating-zone process
Fig. 2-31: Two possible meniscus shapes for a silicon floating zone with
the same values of r
c
=r
f
, L and
M
(11
o
), but different zone volumes and


values. The higher volume mode on the (left) is obviously the more
stable one. From [Teg95].
Spacelab mission D2 showed visible movement with accelerations as
small as 10
-3
-10
-2
g if the disturbance frequencies were close to the zone
resonance frequency (typical values are in the range of 0.5- 10Hz).
The above considerations on the stability of floating zones can in
principle also be applied to free solution zones. In some cases, however,
an important difference related to the wetting between solid and solution
might complicate the situation. In the floating-zone case, the solid-
liquid-gas trijunctions are assumed to be located at the crystal and feed
rod edges, because the wetting angle is usually << 90
o
for a melt in
contact with a solid of the same composition, and the solid-liquid inter-
face position is equal or close to the melting point isotherm. Any
movement away from the edge increases the surface energy of the zone
[Teg95] and is thus unlikely. In the case of a solution zone, the wetting
2. The floating-zone process 65
Fig. 2-32: Calculated maximum zone length L
max
as a function of the
zone radius r for silicon floating zones (
M
=
G
=11
o
) under 1g and for
r
c
/r
f
=1, from [Teg95].
angle of the liquid can be >90
o
, in which case a free zone would detach
from the solid, even under microgravity
c
. If the wetting angle is < 90
o
,
the solution zone can be anchored at the crystal/feed rod edges, but for
systems with low wetting angles it might also creep over it. It must be
kept in mind that the usual measure (in model systems) to prevent fluid
creep in the case of low wetting angles, sharp edges at the solid end
disks, cannot be employed here. Any prefabricated sharp edge is
energetically unstable in contact with a solvent due to its large surface
and will be dissolved. A solution zone with a solvent showing this
behavior is completely unstable under gravity, even with a zone length
well below the static stability limit. Such a zone does not rupture
immediately, but the liquid is slowly drained out of the zone by seeping
over the interface edge and down the side of the crystal. An example is
the growth of GaSb from Bi solution zones, where in the case of THM
growth the strong differences in the thermal expansion coefficients of
GaSb, Bi and ampoule material (SiO
2
) lead to a large amount of stress or
even cracks in the crystal. A floating Bi solution zone is clearly prefer-
able in this case. A calculation of L
max
after equation 2-XVIII for Bi ( =
38010
-3
Nm
-1
,
l
= 10.07 gcm
-3
[CRC81]) leads to a value of 5.2mm, but
experiments with zone lengths of 3.5 and 3mm failed within 1h of
processing time due to Bi seeping over the seed edge [Lau97]. By
processing under microgravity this problem can be overcome, and the
experiment has been performed in May/June 1996 on the Space Shuttle
flight STS-77 in the CFZF mirror furnace (see fig. 2-16); the GaSb
crystal diameter was 15mm and the height of the prefabricated Bi zone
6mm (figs. 2-33 and 2-34). After melting and during the rest of the
66 2. The floating-zone process
c
A liquid not wetting the solid does not preclude a chemical reaction or dissolution;
one important example is liquid silicon in contact with SiO
2
or Al
2
O
3
. For material
combinations with
W
>90
o
, growth with a standard THM arrangement in an ampoule
would therefore be possible.
heat-up time, the zone
seeped over the edges of
the solid rods also under
g and symmetrically
wetted the sides of the
crystal and the feed rod.
This resulted in a much
larger external zone
length than in the case
of THM experiments in
ampoules. The zone
length was also larger
than the external inter-
face distance after
processing (fig. 2-34).
About 2 hours after the
start of heat-up,
however, the zone edges
reached an equilibrium
position and the zone was stable for the rest of the processing time (fig.
2-34 left). The position of the solution zone edges depend on the
temperature. Upon cooling down (but well above the solidification
temperature of bismuth) the external zone height slowly decreased until
it reached the edges of the main interfaces as seen in figs. 2-33 (right)
and 2-34 (right).
2. The floating-zone process 67
Fig. 2-33: Ampoule for the FSZ growth of
GaSb (crystal diameter: 15mm) from a
bismuth solution under g. Left: prefabri-
cated seed, Bi-zone and feed rod assembly
fixed with graphite rings in a fused quartz
ampoule. Right: The same ampoule after
processing. The experiment was performed in
the CFZF (fig. 2-16) on SPACEHAB-4 during
the STS-77 flight (1996) of the space shuttle.
Processing time was 1 day and the transla-
tion rate 1.44mm/d.
2.2.1.2 Transient growth conditions
In addition to the static stability of a floating zone, one must also
consider the mutual dependence of the zone shape and the crystal shape
during growth. A key role in this process is played by the so- called
growth angle
G
, the angle between the meniscus surface and the growth
direction at the circumference of the crystal at the solid-liquid-gas (slg)
trijunction (fig. 2-35). The growth angle is not only important for the FZ
process, but also for CZ and EFG methods. Intuitively, one would expect

G
= 0
o
for most materials, and this has been suggested for different
reasons by several authors [Gau61, And94]. This is, however, incompati-
ble with experimental results showing that in a lot of cases
G
> 0 (table
2-e). One explanation takes into account the differences in the interface
68 2. The floating-zone process
Fig. 2-34: Free Bi solution zone under microgravity, compare also figs.
2-33 and 2-50. Left: Video image of the zone during processing. Center:
Zone and interface shapes derived from the zone pictures and micro-
graphs of the solidified zone; the dark rectangle denotes the size and
position of the solution zone before processing. Right: Axial cut of
crystal, zone, and feed rod after processing. The small wetting angle of
the zone, leading to the spreading of liquid beyond the interface edges
determined by the temperature field and the solubility, is clearly evident.
free energies
lg
,
sl
, and
sg
where the indices lg, sl, and sg denote the
interfaces liquid-gas, solid-liquid, and solid-gas, respectively.
lg
is equal
to the surface tension if adsorption is neglected. Fig. 2-35 shows the
definitions of the various free energies and angles being important for
the determination of the crystal shape.
Bardsley et al. [Bar74] derived a dependence of the growth angle on
the various interface energies from thermodynamic equilibrium consid-
erations (see fig. 2-35 for symbols):
(2-XX)
G
= arccos
o
sg
2
+ o
lg
2
o
sl
2
2$o
sg
$o
lg
2. The floating-zone process 69
Fig. 2-35: Important parameters controlling the floating-zone process:

lg
,
sl
, and
sg
are the interface free energies liquid-gas, solid-liquid
and solid-gas, respectively.
M
is the meniscus angle determined by the
Laplace equation,
G
is the growth angle, the angle
ic
is determined by
the crystal-melt interface curvature, and the crystal angle
c
=
M
-
G
.
Here,
G
is a material constant and the dependence on the growth
geometry, as indicated by measurements of Surek and Chalmers [Sur75,
Sur76a], is neglected. A slightly different approach was taken by Teget-
meier [Teg95]. He did not consider an equilibrium situation, but started
with an arbitrary angle between the seed rod and the zone meniscus, then
assumed a small movement of the zone due to crystallization and calcu-
lated the minimum for the sum of the energy differences of the zone
resulting from this movement. Boundary conditions are a constant
diameter of the feed rod, and a constant zone length. As in eq. 2-XX, the
dependence of
sl
and
sg
on the crystal orientation, as well as the
dependence of all interface energies on impurities, is neglected. The
relation between the interface energies and the different angles is then
given by (see fig. 2-35 for symbols):
The growth angle
G
is thus determined by :
In this case, the growth angle is not a material constant, but is depend-
ent on the meniscus angle (initially or for nonstationary conditions), and,
more important, on the interface curvature. For a flat interface (
i
0
o
)
and stationary growth (
G

M
) eq. 2-XXII is simplified to:
(2-XXIII)
G
= arcsin
o
sl
o
lg
For relatively small values of
i
and
M
, and
sl

lg
eq. 2-XXII can be
linearized:
(2-XXIV)
G
=
o
sl
o
sg
+ 1
o
lg
o
sg
$
(

M

i
)
70 2. The floating-zone process
Eq. 2-XXIV shows that for a growth angle
G
of 0
o
the following
conditions must apply:
sg
=
lg
and
sl
=0. The same conditions for
G
=0
o
result from eq. 2-XX. In accordance with Young's equation for the
wetting angle
W
between a liquid and a solid,
(2-XXV) o
lg
$ cos
(

W
)
= o
sg
o
sl
these two conditions can only be fulfilled simultaneously if
W
=0
o
. This
relation has also been derived by Wenzl et al. [Wen78]. From these
considerations it is obvious that only materials showing complete
wetting by their melt have growth angles close to 0
o
(table 2-e); water
and some metals are examples. Most semiconductors and oxides,
however, show incomplete wetting (
W
is 33
o
for Si and 30
o
for Ge
[Sat80]) and hence
G
>0
o
(fig. 2-36 and table 2-e). A recent investiga-
tion on the solidification of water droplets [And96] also indicates a
possible dependence of
G
on the growth rate. Although
G
is thus
dependent on some external parameters and is not really a material
2. The floating-zone process 71
(2-XXII
$ cos
(

i
)

sin
(

i
)
cos
(

M
)
+arcsin
o
sl
o
sg

o
lg
o
sg
$ sin
(

M
)
$
cos
(

i
)
a

G
=
M
0
c
=
M

i
constant, the variation is small for usual growth conditions (stationary
growth and
i
of the order of a few degrees). For calculating zone and
crystal shapes, the variation can therefore be neglected without introduc-
ing too large an error. Unfortunately, the average growth angle has been
measured only for a few materials, listed in table 2-e. Large growth
angles, like that of InSb or GaSb, indicate a comparatively large value of

sl
(e.g.
sl
= 0.5
lg
for
G
=
M
=30
o
,
ic
=0
o
, eq. 2-XXIII).
One simplification that was made for both eq. 2-XX and eq. 2-XXII is
that the free interface energies
sl
,
sg
, and
lg
do not vary for a given
material system and orientation. Whereas dependences of
sl
,
sg
, and
lg
on temperature or pressure can usually be neglected

for conservative
melt growth systems, this is not true for the composition and the orienta-
tion of the solid. A value of
W
>0
o
already suggests a kinetic energy
barrier for the liquid-solid phase transition [Wen78]. One would there-
fore expect a strong orientation dependence for crystalline materials, due
to the different surface energies of the lattice planes/facets. That this is
indeed the case, has been shown by the results of Dreeben et al. [Dre80]
for Al
2
O
3
, where a doubling was found for
G
when a {0001} facet in
contact with the melt was compared with an atomically rough surface
(table 2-e).
The dependence of
G
on growth kinetics can be expressed by [Tat94]:
72 2. The floating-zone process
Table 2-e (opposite page): Growth angles
G
for different materials;
some of the variations between authors are probably due to different
growth conditions, e.g. interface curvatures, orientations and growth
rates, or to the influence of impurities on interface energies.
* calculated value ** calculated from g drop solidification experi-
ments, data for Si taken from [Kl84, Kl86].
*** Indices in [ ] and denote growth directions, indices in ( ) and { }
the orientation of the facet in contact with the melt meniscus.
2. The floating-zone process 73
[Fei86] 8 100 Y
3
Al
5
O
12
(YAG)
[Fei86] 4 [0001] LiNbO
3
[Dre80] 354 [10!0]
at {0001} facet
[Dre80] 174 [0001]
[Sat80] 122 -
Al
2
O
3
[Lan91d] 2.5 - NaNO
3
[San87, Mar88
And96]
0 - H
2
O
[San86]** 30-33 -
[Sat80] 251 - InSb
[Teg95a] 30.72 111 GaSb
[Teg95a] 17.31.6 100
[Hur63] 15 - GaAs
[Wen78] 01 - Au
[Wen76,Wen78 01 [110] Cu
123 [010] Ga (in 1n HCl solution)
[Wen78] 01 [010] Ga
[Sat80]* 121 isotropic
[Wen78] 73 -
8 polycrystalline
[Sur76a] 13 {111}
Ge
[Teg95a] 8.13 100
[San86]** 11 111
[Sat80]* 111 {111}
[Sur75, Sur76a 11 {111}
Si
Ref.
G
[
o
] Orientation of crystal
or crystal surface***
Material
(2-XXVI)
G
=
G0
k

G
$
AT
s
v
sgl
1
3
where
G0
=growth angle without kinetic effects, k
G
=kinetic coefficient,
T
s
=supercooling of the crystallization front, v
slg
=displacement rate of
the slg trijunction.
Estimated values for T
s
in the case of silicon are 3.7K for
dislocation-free growth of a facet, 0.3K for dislocated growth of a facet
and about 0.05K for the growth of an atomically rough surface [Tat94].
74 2. The floating-zone process
Fig. 2-36: Float zones of several semiconductors in a double ellipsoid
mirror furnace, stationary growth conditions (
M
=
G
, see table 2-e for

G
values). The different appearance of the GaSb zone is due to
additional backlighting of the sample. It can be clearly seen that all
materials have growth angles >0
o
, and that the values of
G
also vary,
with GaSb having the largest value.
The surface energies
lg
and
sg
are also altered by substituting the
gaseous phase (usually vacuum or an inert gas) surrounding the melt
zone and the crystal with an immiscible liquid such as an encapsulant.
The resulting surface energy change is not necessarily the same for the
solid and the liquid phase and may lead to a substantial change of
W
and

G
. One example is Ga, where according to [Wen78]
G
=0
o
in a He
atmosphere, but
G
=12
o
for growth in a 1n HCl solution (table 2-e).
The value of
G
is important for growth situations deviating from
stationary conditions, namely at the start of crystal growth and when a
perturbation occurs, e.g. a change of the zone length (volume) due to a
temperature fluctuation. If the growth system does not allow separate
translations of the crystal and feed rod, e.g. the samples are mounted in
ampoules, the volume and length of the zone cannot be controlled
independently. Then
M

G
(see fig. 2-35), and a positive value of
G
can lead to a breaking of the zone below the stability limits given in
section 2.2.1.1.
In the case of ampoule-mounted samples, the initial float zone is
usually established by melting a cylindrical rod of constant diameter.
Under g, and depending on the volume change upon melting, the zone
will be either curved inward for
l
>
s
, cylindrical for
l
=
s
, or slightly
barrel-shaped for
l
<
s
(compare table 2-c). In most cases, the resulting
meniscus angle
M
will be definitely smaller than the growth angle
G
,

l
>
s
being the worst case because of the negative initial value of
M
.
The material will then crystallize with a smaller diameter than the start-
ing material. Assuming constant zone length, crystallization will thus
lead to a subsequent increase of the zone volume and
M
because of mass
conservation. The reduction of the crystal diameter will therefore be
decreased in the course of the process, until steady state conditions with
a barrel-shaped zone are reached. This can be seen in fig. 2-29 lower left,
2. The floating-zone process 75
fig. 2-38 and, exhibiting the dramatic effect in the case of large growth
angles, in fig. 2-37. Fig. 2-37, showing a recrystallized zone, also illus-
trates the fact that the effect of forming too large a zone cannot be
completely reversed by a subsequent temperature reduction for
G
>0
o
,
because the diameter of the resolidified material will be smaller than that
of the unmelted rod. The evolution of the crystal and zone shape can be
calculated by solving numerically the set of differential equations
describing the change of the zone volume, crystal and feed rod radii, and
meniscus shape over time (see [Mar86, Her94, Teg95]). Fig. 2-38 shows
an example for the crystal growth of a GaAs crystal under g, from the
Spacelab mission D-2 in 1993.
Due to the initial reduction of the crystal diameter with
G
> 0
o
, the
maximum zone length is reduced in comparison with the standard
Rayleigh limit calculated by eq. 2-XVII for the initial zone size. This may
lead to the breaking of the zone and has occurred during one of the first
FZ experiments under g on Spacelab-1 (FSLP) in 1983, where a Si
zone broke during the crystallization of the thin neck part of the crystal
[Eye84a]. A subsequent analysis of the growth process, with zone shapes
taken from video recordings, clearly showed the influence of the growth
angle on the breaking of the zone [Mar86].
Even if it does not lead to the breakage of a zone, the initial diameter
reduction is undesirable because it reduces the volume of the crystal that
can be grown within the - usually quite limited - translation distance. The
effect is of course more pronounced for materials with large values of
G
and/or large volume reductions upon melting, like GaSb or InSb (tables
2-e, 2-c). One can avoid this by starting with a larger zone volume which
results in
M
=
G
from the beginning. This is possible by enlarging the
crystal diameter in the area of the starting material where the first zone is
formed (figs. 2-39 and 2-40). The necessary amount of additional
76 2. The floating-zone process
material can be determined by first deciding on the desired zone length
and then calculating the evolution of the crystal shape for different zone
volumes. Fig. 2-39 shows the different crystal shapes for a GaSb crystal
of 16mm diameter, with a starting material of constant diameter on one
hand, and with an appropriately sized enlargement of the diameter on the
other hand. This approach was successfully tested for GaSb during the
Spacehab-4 mission in 1996; fig. 2-40 shows a sketch of the ampoule
configuration, fig. 2-41 the starting material and the crystals grown from
it.
Under 1g, the situation is different because the meniscus angle is a
function of the zone length. A positive value of
M
is nearly always
obtained, and careful adjustment of the zone length (i.e. the heating
2. The floating-zone process 77
Fig. 2-37: GaSb crystal, grown
under g by recrystallizing a
floating zone. The right insert
shows a video image of the
actual zone. The experiment
was performed during the
sounding rocket campaign
TEXUS 32 within 6 min of g
time, precluding any
translation. The missing 20% of
the zone was still liquid upon
reentry and was hurled to the
ampoule walls by g-forces. The
strongly inward curving shape
of the crystal as well as the
pronounced barrel-shape of the
zone are due to the large
growth angle of this material
(table 2-e).
power) allows a starting condition of
M
=
G
(see e.g. the 1g sample in
fig. 2-41 right). Problems arise, if
M
is increased over
G
by an external
perturbation or by a fluctuation of the control loop. This situation has
been studied by Surek and Coriell [Sur76b, Sur77] using linear perturba-
tion analysis, and by others [Tat94]. They could show that for a stable
system any fluctuation leads to an exponentially damped oscillation of
the crystal diameter around steady state (figs. 2-42, 2-43); for larger
perturbations, the system becomes unstable, the oscillation is no longer
damped but amplified, finally leading to the breakage of the zone. Fig.
2-42 shows both situations for the FZ growth of silicon. The starting
point was a stationary growth situation with a zone length well below the
Heywang limit. The perturbation was a step increase of the power, 6% in
the upper row, 12% in the lower row. The series in the upper row shows
the stable situation with the crystal diameter oscillating around the
78 2. The floating-zone process
Fig. 2-38: Calculated zone and crystal shapes for the FZ-GaAs experi-
ment "FLOBE", performed under g on the Spacelab mission D2 in
1993. The right image shows the crystal obtained together with its calcu-
lated final shape (blue outline). After [Teg95].
starting value, which is given by the feed rod diameter (12mm ) and the
crystal and feed rod translation rates (both 2mm/min); fig. 2-43 shows a
crystal resulting from this type of damped oscillation. The lower series in
fig. 2-42 shows a situation where the stability limit is exceeded, leading
finally to the failure of the zone. It should be emphasized that this break-
age is not due to too high a zone - the zone length is still below the
Heywang limit. The breakage is instead driven by the decrease of the
actual zone volume - caused by the higher amount of crystallized
material due to the larger crystal diameter - in combination with the
2. The floating-zone process 79
Fig. 2-39: Evolution of zone and crystal shapes for GaSb FZ growth
under microgravity, calculated for a zone aspect ratio of 0.8. Top:
standard setup with cylindrical (16mm) starting material. Bottom:
setup with enlarged diameter (20.3mm, 12mm height) of the starting
material in the area of the first melt zone.
different crystal diameter/feed rod diameter ratio and a larger zone
length, requiring a larger volume for a stable zone.
Similar to the g case, a situation with
M
>
G
cannot be completely
reversed by a temperature decrease, but the absolute value of
G
is not as
important here. Under gravity, a large growth angle is actually helpful
with respect to the stability in transient growth situations. Compared to a
80 2. The floating-zone process
Fig. 2-40: Ampoule design used for a GaSb FZ growth experiment
under g in SPACEHAB-4/CFZF (fig. 2-16). A similar setup, but
without the enlarged diameter part, was used for the GaAs-FZ experi-
ment FLOBE on the D2 mission [Cr94b, Cr96], compare fig. 2-38.
The graphite felts simultaneously provide space for thermal expansion
of the crystal before melting and damping of vibrations during launch.
material with the same density and surface tension, but a smaller value of

G
, it permits the use of larger zones (i.e. larger values of the meniscus
angle
M
) for stationary growth

(compare fig. 2-36 right).
2. The floating-zone process 81
Fig. 2-41: GaSb FZ crystals from the Spacehab-4/STS-77 space shuttle
mission May/June 1996. From left to right are shown the unprocessed
flight and flight spare samples FR15 and FR9 in their ampoules, the
same samples after processing, and a 1g sample (see fig. 2-36 right for
the corresponding 1g zone) for comparison. The shape of the processed
g crystals demonstrates the successful result of using enlarged diameter
parts in the starting material. This is especially obvious when compared
to the crystal shape in fig. 2-37. The small initial decrease of the crystal
diameters of FR15 and FR9 indicates a slightly longer first melt zone
than originally planned. All samples are single crystalline (note the
{110} facets in the processed samples, e.g. at the center of FR15). The 1g
sample has the maximum diameter that can be grown on earth by the FZ
process without RF heating.
82 2. The floating-zone process
2. The floating-zone process 83
Fig. 2-43: 111 silicon crystal of 10mm , grown with 2mm/min and
showing the oscillation of the crystal diameter after a step increase of
the power (red arrow mark). The oscillation is completely decayed after
three cycles. Growth direction is from left to right.
Fig. 2-42 (opposite page): Evolution of the zone and crystal shapes of
two silicon float-zones in a mirror furnace after introducing a controlled
perturbation (step increase of the heating power by 6% and 12%, respec-
tively). In both cases, the starting conditions were stationary growth and
a zone length well below the Heywang limit. The 6% increase shown in
the upper row depicts the self-stabilizing behavior of the floating-zone
process, with the crystal diameter performing an exponentially damped
oscillation around steady state. In the lower row, the increase was 12%
and the oscillation of the diameter is not damped but amplified, leading
finally to the breaking of the zone. Even in this case (lower row right),
the zone length was still below the Heywang limit; the failure of the zone
was due to the combination of an enlarged zone length and a reduced
zone volume.
5. References
N.I. Autenshlyus, Yu.P. Betin, V.V. Dobrovensky, A. Ya. Zbarsky, U.I.
Kotika and I.L. Shenderovich: Control system for diameter of crystals
grown from the melt. Priboryi Tekhnika Eksperimenta 2 (1975), 226
Aut75
P.A. Arsenjew, M.N. Branow, K. Bienert and E.F. Kustov: Zur Zchtung
von Einkristallen oxidischer Verbindungen fr die Quantenelektronik
nach der Methode des optischen Zonenschmelzens. Kristall und Technik
(Cryst. Res. Technol.) 8 (1973), 1113
Ars73b
P.A. Arsenjew, M.N. Baranow, E.F. Kustow and K. Bienert: Zur
Zchtung von Einkristallen nach der Methode des optischen
Zonenschmelzens. Experimentelle Technik der Physik 21 (1973), 289
Ars73a
A.V. Anilkumar, R.N. Grugel, X.F. Shen, C.P. Lee and T.G. Wang:
Control of thermocapillary convection in a liquid bridge by vibration. J.
Appl. Phys. 73 (1993), 4165
Ani93
D.M. Anderson, M. Grae Worster, S.H. Davis: The case for a dynamic
contact angle in containerless solidification. J. Crystal Growth 163
(1996), 329
And96
D.M. Anderson and S.H. Davis: Fluid flow, heat transfer and solidifica-
tion near tri-junctions. J. Crystal Growth 142 (1994), 245
And94
C. Allen: The surface tension of liquid metals. In: Liquid Metals,
Chemistry and Physics, Ed. S.Z. Beer (Marcel Dekker, New York1972)
All72
J.I.D. Alexander, S. Amiroudine, J. Ouazzani and F. Rosenberger:
Analysis of the low-gravity tolerance of Bridgman-Stockbarger crystal
growth. II: Transient and periodic accelerations. J. Crystal Growth 113
(1991), 21
Ale91
J.I.D. Alexander, J. Ouazzani and F. Rosenberger: Analysis of the
low-gravity tolerance of Bridgman-Stockbarger crystal growth. I: Steady
and impuls accelerations. J. Crystal Growth 97 (1989), 285
Ale89
T. Alboussire, R. Moreau and D. Camel: Influence dun champ magn-
tique sur la solidification dalliages mtalliques. C.R. Acad. Sci. Paris, t.
313/II (1991), 749
Alb91
T. Akashi, K. Matsumi, T. Okada and T. Mizutani: Preparation of ferrite
single crystals by new floating zone technique. IEEE Trans. Mag. 5
(1969), 285
Aka69
R. Abbaschian, A. Gokhale, E. Jensen and R. Panchapakesan: Liquid
encapsulated molten zone (LEMZ) processing on STS-57. Proc. of the
32nd Conf. of the Amercan Institute of Aeronautics and Astronautics,
Reno, NV 1994 , 563
Abb94
5. References 219
C. Belin: On the growth of large single crystals of calcite by travelling
solvent zone melting. J. Crystal Growth 34 (1976), 341
Bel76
C. Belin, J.J. Brissot and R.E. Jesse: The growth of calcite single crystals
by travelling solvent zone melting. J. Crystal Growth 13/14 (1972), 597
Bel72
R.J. Behrle, H. Figgemeier, A. Danilewsky and K.W. Benz: Zone
melting furnace, development and test of a modular crystal growth facil-
ity for g-applications. Proc. VIIth Europ. Symp. on Materials and Fluid
Sciences in Microgravity , Oxford 1989 (ESA-SP295, 1989), 673
Beh89
J.G. Bednorz and H. Arend: A 1kW mirror furnace for growth of refrac-
tory oxide single crystals by floating-zone technique. J. Crystal Growth
67 (1984), 660
Bed84
J. Baumgartl: Numerische und experimentelle Untersuchungen zur
Wirkung magnetischer Felder in Kristallzchtungsanordnungen. Ph.D.
thesis, University of Erlangen 1992
Bau92
J. Baumgartl, W. Budweiser, G. Mller and G. Neumann: Studies of
buoyancy driven convection in a vertical cylinder with parabolic
temperature profile. J. Crystal Growth 97 (1989), 9
Bau89
E. Bauser and H.P. Strunk: Microscopic growth mechanisms of semicon-
ductors: Experiments and models. J. Crystal Growth 69 (1984), 561
Bau84
F.J. Baum: Radiant energy zone heating unit. Rev. Sci. Instr. 30 (1959),
1064
Bau59
J. Barthel, K. Eichler, M. Jurisch and M. Lser: On the significance of
surface tension driven flow in floating zone melting experiments. Kristall
und Technik 14 (1979), 637
Bar79
W. Bardsley, F.C. Frank, G.W. Green and D.T.J. Hurle: The meniscus in
Czochralski growth. J. Crystal Growth 23 (1974), 341
Bar74
J. Barthel and K. Eichler: ber den Einflu des schichtweisen Einbaus
von Fremdelementen beim Zonenschmelzen auf den effecktiven Vertei-
lungskoeffizienten. Kristall und Technik 2 (1967), 205
Bar67
G. Balakrishnan, D.McK. Paul, M.R. Lees, A.T. Boothroyd: Single
crystal growth of Bi
2
Sr
2
CaCu
2
O
8
using an infrared image furnace.
Physica C 206 (1993), 148
Bal93
A.M. Balbashov and S.K. Egorov: Apparatus for growth of single
crystals of oxide compounds by floating zone melting with radiation
heating. J. Crystal Growth 52 (1981), 498
Bal81
Kh.S. Bagdasarov, V,V. Dyachenko, A.M. Kevorkov and A. Kholov:
Application of laser heating to crystal growth. In: Growth of Crystals,
Vol. 13, Ed. E.I. Givargizov (Consultants Bureau,/Plenum Publishing
Corp., New York 1986), 364
Bag86
220 5. References
C.A. Burrus and J. Stone: Growth of single crystal sapphire-clad ruby
fibers. Appl. Phys. Letters 31 (1977), 383
Bur77
J.A.Burton, R.C. Prim and W.P. Slichter: The distribution of solute in
crystals grown from the melt. J. Chem. Phys. 21 (1953), 1987
Bur53
F.-U. Brckner and K. Schwerdtfeger: Single crystal growth with the
Czochralski method involving rotational electromagnetic stirring of the
melt. J. Crystal Growth 139 (1994), 351
Br94
M. Bruder, A. Crll, A. Eyer, H. Leiste, W. Mller and U. Probst:
Zonenkristallisationsexperimente in Spiegelfen/Zone crystallization
experiments in mirror heating facilities. Forschungsbericht/Final Report
BMFT-FB-W86-023 (ISSN 0170-1339)
Bru86
W.S. Brower and H.S. Parker: Growth of single crystal cuprous oxide. J.
Crystal Growth 8 (1971), 227
Bro71
F. Brown and W.H. Todt: Floating-zone BaTiO
3
: preparation and proper-
ties. J. Appl. Phys. 35 (1964), 1594
Bro64
J.D. Broder and G.A. Wolff: A new method of GaP growth. J. Electro-
chem. Soc. 110 (1963), 1150
Bro63
J.C. Brice: The growth of crystals from liquids. (North-Holland, Amster-
dam 1973)
Bri73
J.J. Brissot and C. Belin: Preparation of artificial calcite single crystals
by solvent zone melting. J Crystal Growth 8 (1971), 213
Bri71
E.A. Boucher and M.J.B. Evans: A liquid-bridge model for the float-zone
processing of materials. J. Chem. Soc., Faraday Trans.1 81 (1985), 2787
Bou85
J. Bohm, A. Ldge and W. Schrder: Crystal growth by floating zone
melting. In: Handbook of Crystal Growth 2a, Ed. D.T.J. Hurle
(Elsevier/North-Holland, Amsterdam 1994), 213
Boh94
R.V. Birikh, V.A. Briskman, V.I. Chernatynsky and B. Roux: Control of
thermocapillary convection in a liquid bridge by high frequency vibra-
tions. Microgravity Q. 3 (1993), 23
Bir93
J. P. Birat and J. Chon: Electromagnetic stirring on billet, bloom, and
slab continuous casters: State of the art in 1983. Ironmaking and Steel-
making 10 (1983), 269
Bir83
K.W. Benz: Herstellung und Eigenschaften des Werkstoffes GaSb fr
Anwendungen in der Optoelektronik. Habilitation thesis (University of
Duisburg 1982)
Ben82
K.W. Benz and E. Bauser: Growth of binary III-V semiconductors from
metallic solutions. In: Crystals - Growth, Properties and Applications,
Ed. H.C. Freyhardt (Springer, Berlin 1980),1
Ben80
K.W. Benz and G. Mller: GaSb and InSb crystals grown by vertical and
horizontal travelling heater method. J. Crystal Growth 46 (1979), 35
Ben79
5. References 221
G.K. Celler, L.E. Trimble and Lynn O.Wilson: Kinetics of radiative
melting of Si. In: Proc. Energy Beam-Solid Interactions and Transient
Thermal Processing Symp., Boston 1984, Eds. D.K. Biegelsen, G.A.
Rozgonyi, C.V. Shank (Mat. Res. Symp. Proc. Vol. 35 1985), 635
Cel85
G.K. Celler, K.A. Jackson, L.E. Trimble, McD. Robinson and D.J.
Lischner: Faceted melting and superheating of crystalline Si irradiated
with incoherent light. In: Proc. Energy Beam-Solid Interactions and
Transient Thermal Processing Symp., Boston 1983, Eds. J.C.C. Fan and
N.M. Johnson (North-Holland, New York 1984), 409
Cel84
T. Carlberg and M. Levenstam: Laser-heating applied to microgravity
experiments - a feasibility study. Microgravity Sci. Technol. IV/4 (1991),
254
Car91
T. Carlberg: Some aspects on the formation of striations during crystal
growth from the melt. J. Crystal Growth 85 (1987), 32
Car87
T. Carlberg: Lateral solute segregation during floating-zone crystal
growth under different gravity conditions. J. Crystal Growth 79 (1986),
71
Car86b
T. Carlberg: Floating zone experiments with germanium crystals in
sounding rockets. Acta Astronautica 13 (1986), 639
Car86a
T. Carlberg: A preliminary report on floating-zone experiments with
germanium crystals in a sounding rocket. In: Proc. 5th European Symp.
on Materials Sciences und Microgravity, Schlo Elmau 1984
(ESA-SP222), 367
Car84
J.R. Carruthers: Thermal convection instabilities relevant to crystal
growth from liquids. In: Preparation and Properties of Solid-State
Materials Vol.3, Ed. W.R. Wilcox. (Marcel Dekker, New York 1977), 1
Car77
J. R. Carruthers and A.F. Witt: Transient segregation effects in Czochral-
ski growth. In: Crystal growth and Characterization, Eds. R. Ueda and
J.B. Mullin (North-Holland, Amsterdam 1975), 107
Car75
D. Camel and P. Tison: Semi-confined Bridgman growth of Ga-doped
germanium in Maser-2. Proc. VIIth Europ. Symp. on Materials and Fluid
Sciences in Microgravity, Oxford 1989, 63
Cam89
D. Camel and J.J. Favier: Thermal convection and longitudinal
macrosegregation in horizontal Bridgman crystal growth. II. Practical
laws. J. Crystal Growth 67 (1984), 57
Cam84b
D. Camel and J.J. Favier: Thermal convection and longitudinal
macrosegregation in horizontal Bridgman crystal growth. I. Order of
magnitude analysis. J. Crystal Growth 67 (1984), 42
Cam84a
A. Calverley, M. Davis and R.F. Lever: The floating-zone melting of
refractory metals by electron bombardment. J. Sci. Instr. 34 (1957), 142
Cal57
222 5. References
S.R. Coriell and R.F. Sekerka: Lateral solute segregation during unidi-
rectional solidification of a binary alloy with a curved solid-liquid inter-
face I. J. Crystal Growth 46 (1979),479
Cor79
S.R. Coriell, S.C. Hardy and M.R. Cordes: Stability of liquid zones. J. of
Colloid and Interface Sciences 60 (1977), 126
Cor77b
S.R. Coriell and M.R. Cordes: Theory of molten zone shape and stability.
J. Crystal Growth 42 (1977), 466
Cor77a
P.A. Clark and W.R. Wilcox: Influence of gravity on thermocapillary
convection in floating zone melting of silicon. J. Crystal Growth 50
(1980), 461
Cla80
W. Class: Growth of yttrium aluminate and yttrium aluminium garnet by
a hollow cathode floating-zone method. J. Crystal Growth 3/4 (1968),
241
Cla68
W. Class, H.R. Nesor and G.T. Murray: Preparation of oxide crystals by
a plasma float-zone technique. In: Crystal Growth, Ed. H.S. Peiser
(Pergamon Press, Oxford 1967), 75
Cla67
C.H. Chun and H. Wuest: Thermische Marangonikonvektion in einer
Schwebezone: g-Experiment whrend des Raketenflugs von TEXUS
IIIb. Z. Flugwiss. Weltraumforsch. 6 (1982), 316
Chu82
C.H. Chun: Marangoni convection in a floating zone under reduced
gravity. J. Crystal Growth 48 (1980), 600
Chu80
C.H. Chun and H. Wuest: Experiments on the transition from the steady
to the oscillatory Marangoni convection of a floating zone under reduced
gravity effect. Acta Astronautica 6 (1979), 1073
Chu79
C.H. Chun and H. Wuest: A microgravity simulation of the Marangoni
convection. Acta Astronautica 5 (1978), 681
Chu78
J.-C. Chen and C. Hu: Measurement of the surface temperature distribu-
tion in the float zone of LiNbO
3
. J. Crystal Growth 158 (1996), 289
Che96
J.-C. Chen and C. Hu: Measurement of float-zone interface shape for
lithium niobate. J. Crystal Growth 149 (1995), 87
Che95
H.A. Chedzey and D.T.J. Hurle: Avoidance of growth striae in semicon-
ductor and metal crystals grown by zone-melting techniques. Nature 210
(1966), 933
Che66
C.E. Chang, W.R. Wilcox and R.A. Lefever: Thermocapillary convection
in floating zone melting: influence of zone geometry and Prandtl number
at zero gravity. Mat. Res. Bull. 14 (1979), 527
Cha79
C.E. Chang and W.R. Wilcox: Analysis of surface tension driven flow in
floating zone melting. Int. J. Heat Mass Transfer 19 (1976), 355
Cha76
C.E. Chang and W.R. Wilcox: Inhomogeneities due to thermocapillary
flow in floating zone melting. J. Crystal Growth 28 (1975), 8
Cha75
5. References 223
A. N. Danilewsky, G. Nagel and K.W. Benz: Growth of GaAs from Ga
solution under reduced gravity during the D2-mission. Cryst. Res.
Technol. 29 (1994), 171
Dan94
A.N. Danilewsky, K.W. Benz and T. Nishinaga : Growth kinetics in
earth- and space-grown InP and GaSb crystals. J. Crystal Growth 99
(1990), 1281
Dan90b
A. N. Danilewsky and K.W.Benz: Crystal Growth in mirror heaters: time
markers by lamp-pulses. J. Crystal Growth 106 (1990), 273
Dan90a
A. Crll, A. Danilewsky, M. Schweizer, A.Tegetmeier and K.W. Benz:
Floating-zone growth of GaAs. J. Crystal Growth 166 (1996), 239
Cr96
A. Crll, A. Tegetmeier, G. Nagel and K.W. Benz: Floating-zone growth
of GaAs under microgravity during the D2-mission. Cryst. Res. Technol.
29 (1994), 335
Cr94b
Segregation in Si floating-zone crystals grown under microgravity and in
a magnetic field. J. Crystal Growth 137 (1994), 95
Cr94a
A. Crll, W. Mller-Sebert, K.W. Benz and R. Nitsche: Natural and
thermocapillary convection in partially confined silicon melt zones.
Microgravity Sci. Technol. III/4 (1991), 204
Cr91
A. Crll and W.Mller-Sebert: Der Einflu von Auftriebs- und Marango-
nikonvektion auf die Segregation bei der Kristallzchtung von Silicium
nach dem Floating-Zone-Verfahren. Ph.D. thesis, University of Freiburg
1988.
Cr88
A. Crll, W. Mller and R. Nitsche: Dopant distribution in semiconduc-
tor crystals under microgravity. In: Proc. 6th Europ. Symp. on Material
Sciences under Microgravity Conditions, Bordeaux 1986 (ESA-SP256
,1987), 87
Cr87
A. Crll, W. Mller and R. Nitsche: Floating-zone growth of surface-
coated silicon under microgravity. J. Crystal Growth 79 (1986), 65
Cr86
R.K. Crouch, A.L. Fripp and W.J. Debnam: Thermophysical properties
of germanium for thermal analysis of growth from the melt. In: Materials
Processing in the Reduced Gravity Environment of Space, Ed. G.E.
Rindone (Elsevier, Amsterdam 1982), 657
Cro82
R.C. Weast and M.J. Astle (Eds.):CRC Handbook of Chemistry and
Physics. 62nd Ed. (CRC Press, Boca Raton, FL 1981)
CRC81
R.T. Cox, A. Revcolevschi and R. Collongues: Growth of an O
17
enriched Al
2
O
3
crystal by a floating zone technique. J. Crystal Growth 15
(1972), 301
Cox72
S.R. Coriell, R.F. Boisvert, R.G. Rehm and R.F. Sekerka: Lateral solute
segregation during unidirectional solidification of a binary alloy with a
curved solid-liquid interface. II Large departures from planarity. J.
Crystal Growth 54 (1981),167
Cor81
224 5. References
R.F. Dressler and N.S. Sivakumaran: Non-contaminating method to
reduce Marangoni convection in microgravity floating zones. J. Crystal
Growth 88 (1988), 148
Dre88
A.B. Dreeben, K.M. Kim and A. Schujko: Measurement of meniscus
angle in laser heated float zone growth of constant diameter sapphire
crystals. J. Crystal Growth 50 (1980), 126
Dre80
J. Dorner: Einflu eines berlagerten axialen Gleichstromes beim Zonen-
ziehen von Siliziumstben auf deren Eigenschaften. Int. Z. fr
Elektrowrme 22 (1964), 331
Dor64
P. Dold and K.W. Benz: Modification of fluid flow and heat transport in
vertical Bridgman configurations by rotating magnetic fields. Cryst. Res.
Technol., accepted for publication 1996/1997
Dol96
P. Dold and K.W. Benz: Convective temperature fluctuations in liquid
gallium in dependence on static and rotating magnetic fields. Cryst. Res.
Technol. 30 (1995), 1135
Dol95
P. Dold: Einflsse statischer und dynamischer Magnetfelder bei der
Kristallzchtung aus Metall- und Halbleiterschmelzen. Ph. D. thesis
Universitt Freiburg i. Br. 1994
Dol94
P.A. Davidson: Magnetic damping of jets and vortices. J. Fluid Mech.
299 (1995), 153
Dav95
M.C. Davidson and L.R. Holland: Narrow zone heating by a new radia-
tion focusing technique - toroidal ellipsoid furnace. Rev. Sci. Instrum. 49
(1978), 1156
Dav78
M.Y. Dashevskii, G.V. Kukuladze, V.B. Lazarev and M.S. Mirgalovskii:
ONCFPUMNQRM[F _CKFMI_ C PAQOKACAU AMRILNMIEA
DAKKI_. Neorg. Mater. 3 (1967), 1561
Das61
W.C. Dash: Improvements on the pedestal method of growing silicon and
germanium crystals. J. Appl. Phys. 31 (1960), 736
Das60
W.C. Dash: Growth of silicon crystals free from dislocations. J. Appl.
Phys. 30 (1959), 459
Das59
W.C Dash: Evidence of dislocation jogs in deformed silicon. J. Appl.
Phys. 29 (1958), 705
Das58
A.N. Danilewsky, S. Lauer, J. Meinhardt, K.W. Benz, B. Kaufmann, R.
Hoffmann and A. Drnen: Growth and characterization of GaSb bulk
crystals with low acceptor concentration. J. Electronic Materials 25
(1996), 1082
Dan96a
A.N. Danilewsky, S. Lauer, G. Bischopink and K.W. Benz: Long term
crystal growth under microgravity during the EURECA-1 mission (I)
THM growth of Al
x
Ga
1-x
Sb. Cryst. Res. Technol. 31 (1996), 11
Dan96
5. References 225
J.J. Favier: Macrosegregation-II: A comparative study of theories. Acta.
Met. 29 (1981), 205
Fav81b
J.J. Favier: Macrosegregation-I: Unified analysis during non-steady state
solidification. Acta. Met. 29 (1981), 197
Fav81a
A. Eyer, H.Leiste and R. Nitsche: Floating-zone growth of silicon in a
sounding rocket. J. Crystal Growth 71 (1985), 173
Eye85b
A. Eyer and H. Leiste: Striation-free silicon crystals by float-zoning with
surface-coated melt. J. Crystal Growth 71 (1985), 249
Eye85a
A. Eyer, H. Leiste and R. Nitsche: Floating-zone growth of silicon in
Spacelab-1. In: Proc. 5th European Symp. on Materials Sciences und
Microgravity, Schlo Elmau 1984 (ESA-SP222), 173
Eye84a
A. Eyer, H. Leiste, M. Schuhmacher and H. Walcher: Entwicklung von
Kristallisationsexperimenten fr Spacelab: Silizium (E1), CdTe (D1),
ZnS (D1)/Bau und Erprobung einer Monoellipsoidspiegelheizanlage -
Phase 2// Preparation of crystal growth experiments in Spacelab: Silicon
(E1), CdTe(D1), ZnS(D1)/ Construction and test of a single ellipsoid
mirror heating facility - phase 2. Forschungsbericht/Final Report BMFT-
FB-W84-045 (ISSN 0170-1339)
Eye84
A. Eyer, H. Leiste, M. Schuhmacher and H. Walcher: Entwicklung von
Kristallisationsexperimenten fr Spacelab: Silizium (E1), CdTe (D1),
ZnS (D1)/Bau und Erprobung einer Monoellipsoidspiegelheizanlage//
Preparation of crystal growth experiments in Spacelab: Silicon (E1),
CdTe(D1), ZnS(D1)/ Construction and test of a single ellipsoid mirror
heating facility. Forschungsbericht/Final Report BMFT-FB- W81-023
(ISSN 0170-1339)
Eye81
A. Eyer, H. Zimmermann and R. Nitsche: A radiation furnace for zone-
crystallization experiments in Spacelab. J. Crystal Growth 47 (1979), 219
Eye79
A. Eyer and H. Zimmermann: Flssig- und Gaszonenkristallisation unter
Schwerelosigkeit/Zone crystallization under zero gravity.
Abschlubericht/Final Report BMFT-FB W 77-12 (1977)
Eye77
Tiegelfreies Ziehen von Siliciumeinkristallen. Z. Naturforsch. 9a (1954),
67
Eme54
D. Elwell, W.L. Kway and R.S. Feigelson: Crystal growth of a new
tetragonal phase of ScTaO
4
. J. Crystal Growth 71 (1985), 237
Elw85
K. Eickhoff and K. Grs: Tiegelfreies Zonenschmelzen von Rubinkristal-
len durch Aufheizen der Schmelzzone mittels Laser. J. Crystal Growth 6
(1969), 21
Eic69
J.L. Duranceau and R.A. Brown: Thermal-capillary analysis of small-
scale floating zones: Steady-state calculations. J. Crystal Growth 75
(1986), 367
Dur86
226 5. References
D. Gazit, P.N. Peszkin, L.V. Moulton and R.S. Feigelson: Influence of
growth rate on the structure and composition of float-zone grown
Bi
2
Sr
2
CaCu
2
O
8
superconductor fibers. J. Crystal Growth 98 (1989), 545
Gaz89
D. Gazit and R.S. Feigelson: Laser-heated pedestal growth of high T
c
Bi-Sr-Ca-Cu-O superconducting fibers. J. Crystal Growth 91 (1988), 318
Gaz88
G.K. Gaule and R.J. Pastore: The role of surface tension in pulling single
crystals of controlled dimensions. In: Metallurgy of Elemental and
Compound Semiconductors, Ed. R.G. Grubel. (Interscience, New York
1961), 201
Gau61
D.B. Gasson and B. Cockayne: Oxide crystal growth using gas lasers. J.
Mat Sci. 5 (1970), 100
Gas70
D.B. Gasson: The preparation of calcium tungstate crystals by a modified
floating zone recrystallization technique. J. Sci. Instrum. 42 (1965), 114
Gas65
J.P. Garandet, S. Corre, S. Gavoille, J.J. Favier and J.I.D. Alexander: On
the effect of gravity perturbations on composition profiles during Bridg-
man crystal growth in space. J. Crystal Growth 165 (1996), 471
Gar96
J.P. Garandet, J.J. Favier and D. Camel: Segregation phenomena in
crystal growth from the melt. In: Handbook of Crystal Growth 2b, Ed.
D.T.J. Hurle (Elsevier/North-Holland, Amsterdam 1994), 659
Gar94
J.P. Garandet, T. Duffar and J.J. Favier: On the scaling analysis of the
solute boundary layer in idealized growth configurations. J. Crystal
Growth 106 (1990), 437
Gar90
W.W. Fowlis and G.O. Roberts: Confinement of thermocapillary floating
zone flow by uniform rotation. J. Crystal Growth 74 (1986), 301
Fow86
W.G. Field, R.W. Wagner: Thermal imaging for single crystal growth
and its application to ruby. J. Crystal Growth 3-4 (1968), 799
Fie68
M.M. Fejer, J.L. Nightingale, G.A. Magel and R.L. Byer: Laser-heated
miniature pedestal growth apparatus for single-crystal optical fibers. Rev.
Sci. Instrum. 55 (1984), 1791
Fej84
R.S. Feigelson: Growth of single crystal fibers. MRS Bulletin XIII/10
(1988), 47
Fei88
R.S. Feigelson: Pulling of optical fibers. J. Crystal Growth 79 (1986),
669
Fei86
J.J. Favier, J.P. Garandet, A. Rouzaud and D. Camel: J. Crystal Growth
140 (1994), 237
Fav94
J.J. Favier: Striations of kinetic origin induced during non-steady-state
solidification. J. Electrochem. Soc. 129 (1982), 2355
Fav82a
J.J. Favier and L.O. Wilson: A test of the boundary layer model in
unsteady Czochralski growth. J. Crystal Growth 58 (1982), 103
Fav82
5. References 227
R. N. Grugel, X.F. Shen, A.V. Anilkumar and T.G. Wang: The influence
of vibration on microstructural uniformity during floating-zone crystal
growth. J. Crystal Growth 142 (1994), 209
Gru94
L.A. Gorbunov and E.D. Lumkis: Thermoelectromagnetic convection in
bulk single crystal growth under weightlessness. In: Proc. of the 1st Int.
Symp. on Hydromechanics and Heat/Mass transfer in microgravity,
Perm-Moscow1991, V.A. Briskman et al. , Eds. (Gordon and Breach,
Reading 1992), 449
Gor92
H. Gonzalez and A. Castellanos: The effect of residual axial gravity on
the stability of liquid columns subjected to electric fields. J. Fluid. Mech.
249 (1993), 185
Gon93
H. Gonzalez, F.M.J. McCluskey, A. Castellanos and A. Barrero: Stabili-
zation of dielectric liquid bridges by electric fields in the absence of
gravity. J. Fluid. Mech. 206 (1989), 545
Gon89
V.G. Glebovsky, V.N. Semenov and V.V. Lomeyko: The characteristic
features of growth and the real structure of tungsten tube crystals. J.
Crystal Growth 98 (1989), 487
Gle89
V.M. Glazov, K. Dovletov, A. Ya. Nashelskii and M.M. Mamedov:
Igkelhe Qqoriqroz Bjhfheqm Omo~di` P`pnj`b` InP Ooh M`qoeb`lhh.
Neorg. Mater. 13 (1977), 34
Gla77
V.M. Glazov, S.N. Chizhevskaya and N.N. Glagoleva: Liquid Semicon-
ductors. (Plenum Press, New York 1969)
Gla69
G.Z. Gershuni, T.V. Lyubimova, D.V. Lyubimov and B. Roux: Coupled
thermovibrational and thermocapillary convection in liquid bridge (float-
ing zone system). Proc. VIIIth Europ. Symp. on Materials and Fluid
Sciences in Microgravity, Bruxelles 1992 (ESA-SP333), 117
Ger92
R. Geray: Kristallzchtung und magnetische Eigenschaften von Kobalt-
und Nickelsilikaten. Ph.D. thesis Universitt Marburg/Lahn 1990
Ger90
R. Geray: Zonenschmelzen von Gadolinium-Gallium Granat in einem
Doppelellipsoidspiegelofen mit Halogenlampenbeheizung. Masters thesis
Universitt Freiburg i. Br. 1984
Ger84
P. Gerthsen: Ein neues Verfahren zur Kristallzchtung von hochschmel-
zenden Stoffen. Z. angew. Physik 15 (1963), 301
Ger63
Y. M. Gelfgat, L.A. Gorbunov and M.Z. Sorkin: Electromagnetic
methods for influencing the hydrodynamics and heat and mass transfer
during growth of large single crystals. Growth of Crystals 16 (1991), 265
Gel91
Y. M. Gel'fgat and L.A. Gorbunow: An additional source of forced
convection in semiconductor melts during single-crystal growth in
magnetic fields. Sov. Phys. Dokl. 35 (1989), 470
Gel89
228 5. References
W.C. Hsieh and C.W. Lan: Experimental study on floating-zone tube
growth and comparison with computer simulation. J. Crystal Growth 165
(1996), 447
Hsi96
M. Higuchi, R.F. Geray, R. Dieckmann, D.G. Park, J.M. Burlitch, D.B.
Barber, C.R. Pollock: Growth of Cr
4+
-rich, chromium-doped forsterite
single crystals by the floating zone method. J. Crystal Growth 148
(1995), 140
Hig95
W. Heywang: Zur Stabilitt senkrechter Schmelzzonen. Z. Naturforsch
11a (1956), 238
Hey56
F.M. Herrmann and G. Mller: Growth of 20mm diameter GaAs crystals
by the floating-zone technique with controlled As-vapour pressurer under
microgravity. J. Crystal Growth 156 (1995), 350
Her95
F.-M. Herrmann: Untersuchungen zum tiegelfreien Zonenschmelzver-
fahren bei der Zchtung von GaAs unter Mikrogravitation. Ph.D. thesis
Universitt Erlangen-Nrnberg 1994
Her94
R.M. Henson and A.J. Pointon: Growth of single crystal Ba
0.65
Sr
0.35
TiO
3
by solvent zone melting. J. Crystal Growth 26 (1974), 174
Hen74
R. Haya, D. Rivas and J. Sanz: Radiative exchange between a cylindrical
crystal and a monoellipsoidal mirror furnace. Int J. Heat Mass Transfer
40 (1996), 323
Hay96
S. Hayashi, and H. Komatsu: Growth of cobalt single crystals by the
electron beam floating-zone method. Krist. u. Techn. 13 (1978), 263
Hay78
K. Hatta, M. Higuchi, J. Takahashi and K. Kodaira: Floating zone growth
and characterization of aluminum-doped rutile single crystals. J. Crystal
Growth 163 (1996), 279
Hat96
I. Harter, P. Dusserre, T. Duffar, J.-Ph. Nabot and N. Eustathopoulos:
Wetting of III-V melts on crucible materials. J. Crystal Growth 131
(1993), 157
Har93
G.M. Harriott and R.A. Brown: Steady solute fields induced by differen-
tial rotation in a small floating zone. J. Crystal Growth 69 (1984), 589
Har84a
S.C. Hardy: The surface tension of liquid silicon. J. Crystal Growth 69
(1984), 456
Har84
R.W. Gurtler, A. Baghdadi, R.J. Ellis and J.A. Lesk: Silicon ribbon
growth via the ribbon-to-ribbon (RTR) technique: Process update and
material characterization. J. Electron. Mater. 7 (1978), 441
Gur78
K.P. Gupta, R.O. Gregory and M. Rossnick: Limitations in using
kilohertz radio frequencies for float zone silicon crystals. J. Crystal
Growth 44 (1978), 526
Gup78
5. References 229
T. Kaiser: Computersimulationen von Strmungsvorgngen in Halbleit-
erschmelzen mit und ohne Magnetfeld. Masters thesis Universitt
Freiburg 1993
Kai93
S. Kaddeche, H. Ben Hadid, D. Henry: Macrosegregation and convection
in the horizontal Bridgman configuration. II: Concentrated alloys. J.
Crystal Growth 141 (1994), 279
Kad94
M. Jurisch: Surface temperature oscillations of a floating zone resulting
from oscillatory thermocapillary convection. J. Crystal Growth 102
(1990), 223
Jur90b
M. Jurisch and W. Lser: Analysis of periodic non-rotational W stria-
tions in Mo single crystals due to nonsteady thermocapillary convection.
J. Crystal Growth 102 (1990), 214
Jur90a
M. Jurisch, W. Lser, E. Lyumkis, E. Martuza ne and B. Martuza ns:
Connection of the thermocapillary flow characteristics and the impurity
distribution pattern in floating zone molten molybdenum crystals. Cryst.
Res. Technol. 17 (1982), 963
Jur82
T. Jung and G. Mller: Effective segregation coefficients: a comparison
of axial solute distributions predicted by analytical boundary layer
models and numerical calculations. J. Crystal Growth 165 (1996), 463
Jun96
D.W. Jones: Refractory metal crystal growth techniques. In: Crystal
Growth, Theory and Techniques, Ed. C.H.L. Goodman (Plenum Press,
London 1974), 233
Jon74
K.A. Jackson and D.A. Kurtze: Instability in radiatively melted silicon
films. J. Crystal Growth 71 (1985), 385
Jac85
T. Imai, S. Yagi, Y. Sugiyama and I. Hatakeyama: Growth of potassium
tantalate niobate single crystal fibers by the laser-heated pedestal growth
method assisted by a crystal cooling technique. J. Crystal Growth 147
(1995), 350
Ima95
D.T.J. Hurle and R.W. Series: Use of a magnetic field in melt growth. In:
Handbook of Crystal Growth 2a, Ed. D.T.J. Hurle (Elsevier/North-
Holland, Amsterdam 1994), 259
Hur94
D.T. J. Hurle and E. Jakeman: Effects of fluctuations on the measure-
ment of distribution coefficients by directional solidification. J. Crystal
Growth 5 (1969), 227
Hur69
D.T.J. Hurle: Mechanisms of growth of metal single crystals from the
melt. Prog. Mat. Sci. 10 (1963), 81
Hur63
H.E. Huppert and J.S. Turner: Double-diffusive convection. J. Fluid.
Mech. 106 (1981), 299
Hup81
X. Huang, S. Togawa, S.-I. Chung, K. Terashima and S. Kimura: Surface
tension of a Si melt: influence of oxygen partial pressure. J. Crystal
Growth 156 (1995), 52
Hua95
230 5. References
K.M.Kim, A.F. Witt, M. Lichtensteiger and H.C. Gatos: Quantitative
analysis of the effects of dwstabilizing vertical thermal gradients on
crystal growth and segregation: Ga-doped Ge. J. Electrochem. Soc. 125
(1978), 475
Kim78a
S. Kimura, I. Shindo, K. Kitamura, Y. Mori and H. Tamakizawa: Evalua-
tion of yttrium iron garnet single crystals grown by the floating zone
method. J. Crystal Growth 44 (1978), 621
Kim78
S. Kimura and I. Shindo: Single crystal growth of YIG by the floating
zone method. J. Crystal Growth 41 (1977), 192
Kim77
K.M. Kim, A.F. Witt and H.C. Gatos: Grystal growth from the melt
under destabilizing thermal gradients. J. Electrochem. Soc. 119 (1972),
1218
Kim72
W. Keller and A. Mhlbauer: Floating-zone silicon. Vol. 5 of: Prepara-
tions and properties of solid-state materials, Ed. W.R. Wilcox (Marcel
Dekker, NewYork 1981)
Kel81
W. Keller: Experimental influence of some growth parameters upon the
shape of the melt interfaces and the radial phosphorus distribution during
float-zone growth of silicon single crystals. J. Crystal Growth 36 (1976),
215
Kel76
P.H. Keck and W. van Horn: The surface tension of liquid silicon and
germanium. Phys. Rev. 91 (1953), 512
Kec53a
P.J. Keck and M.J.E. Golay: Crystallization of silicon from a floating
liquid zone. Phys. Rev. 89 (1953), 1297
Kec53
N.D. Kazarinoff and J.S. Wilkowski: A numerical study of Marangoni
flows in zone-refined silicon crystals. Phys. Fluids A 1 (1989), 625
Kaz89
H.H. Karow: Fabrication methods for precision optics. (J. Wiley, New
York 1993)
Kar93
P. Kapitza: The study of the specific resistance of bismuth crystals and
its change in strong magnetic fields and some allied problems. Proc.
Royal. Soc. London A 119 (1928), 358
Kap28
Y. Kamotani and J. Kim: Effect of zone rotation on oscillatory thermo-
capillary flow in simulated floating zones. J. Crystal Growth 87 (9188),
62
Kam88
Y. Kamotani, S. Ostrach and M. Vargas: Oscillatory thermocapillary
convection in a simulated floating-zone configuration. J. Crystal Growth
66 (1984), 83
Kam84
T. Kaiser, P. Dold, A. Crll and K.W. Benz: FZ growth of Si in magnetic
fields III: Numerical simulations. J. Crystal Growth, to be submitted for
publication
Kai97
5. References 231
C. Kooy and H.J.M. Couwenberg: Zonenschmelzen von Oxiden im
Kohlebogen-Strahlungsofen. Philips Techn. Rundschau 23 (1961/62),
143
Koo61
H. Klker: Crystallization of a silicon sphere. In: Proc. of the Norderney
Symp. on scientific results of the german Spacelab mission D1,
Norderney 1986 (DFVLR-WPF D1, Kln 1986), 264
Kl86
H. Klker: Crystallization of a silicon sphere. In: Proc.5th European
Symp. on Materials Sciences und Microgravity, Schlo Elmau 1984
(ESA-SP222, 1984), 169
Kl84
H. Klker: The behaviour of nonrotational striation in silicon. J. Crystal
Growth 50 (1980), 852
Kl80
H. Kodera: Diffusion coefficients of impurities in silicon melt. Jap. J.
Appl. Phys. 2 (1963), 212
Kod63
N. Kobayashi and W.R. Wilcox: Computational studies of convection
due to rotation in a cylindrical floating zone. J. Crystal Growth 59
(1982), 616
Kob82
N. Kobayashi: Power required to form a floating zone. J. Crystal Growth
43 (1978), 417
Kob78
J. Klber and R. Kiehne: Surface tension of GaAs melt. Cryst. Res.
Technol. 25 (1990), K261
Kl90
K. Kitamura, S. Kimura and K. Watanabe: Control of interface shape by
using heat reservoir in FZ growth with infrared radiation convergence
type heater. J. Crystal Growth 57 (1982), 475
Kit82
K. Kitamura, N.II, I. Shindo and S. Kimura: Interface shape and horizon-
tal variations of Al and Ga contents in substituted YIG single crystals
grown by the floating zone method. J. Crystal Growth 46 (1979), 277
Kit79
K. Kitazawa, N. Nagashima, T. Mizutani, K. Fueki and T. Mukaibo: A
new thermal imaging system utilizing a Xenon arc lamp and an ellipsoi-
dal mirror for crystallization of refractory oxides. J. Crystal Growth 39
(1977), 211
Kit77
M. Kimura, H. Arai, T. Mori and H. Yamagishi: Facet formation in
silicon single crystals grown by VMFZ method. J. Crystal Growth 128
(1993), 282
Kim93
H. Kimura, M.F. Harvey, D.J. O'Connor, G.D. Robertson and G.C.
Valley: Magnetic field effects on floating Si zone crystal growth. J.
Crystal Growth 62 (1983), 523
Kim83
K.M. Kim, A.B. Dreeben and A. Schujko: Maximum stable zone length
in float-zone growth of small diameter sapphire and silicon crystals. J.
Appl. Phys. 50 (1979), 4472
Kim79
232 5. References
C.W. Lan and S. Kou: Effect of rotation on radial dopant segregation in
microgravity floating-zone crystal growth. J. Crystal Growth 133 (1993),
309
Lan93b
C. W. Lan and S. Kou: Radial dopant segregation in zero-gravity
floating-zone crystal growth. J. Crystal Growth 132 (1993), 578
Lan93a
C.W. Lan and S. Kou:Shortened floating zone crystal growth under
normal gravity. J. Crystal Growth 119 (1992), 281
Lan92b
C.W. Lan and S. Kou:A simple method for improving the stability of
float zones under normal gravity. J. Crystal Growth 118(1992), 151
Lan92a
C.W. Lan and S. Kou: Effects of rotation on heat transfer, fluid flow and
interfaces in normal gravity floating-zone crystal growth. J. Crystal
Growth 114 (1991), 517
Lan91e
C.W. Lan and S.Kou: Heat transfer, fluid flow and interface shapes in
floating-zone crystal growth. J. Crystal Growth 108 (1991), 351
Lan91d
C.W. Lan and S. Kou: Floating-zone crystal growth with a heated and
immersed shaper - experiments. J. Crystal Growth 108 (1991), 541
Lan91c
C.W. Lan and S. Kou: Floating-zone crystal growth with a heated and
immersed shaper - computer simulation. J. Crystal Growth 108 (1991),
340
Lan91b
C.W. Lan and S. Kou: Floating-zone crystal growth with a heated ring
covering the melt surface. J. Crystal Growth 108 (1991), 1
Lan91a
C.W. Lan and S. Kou: Thermocapillary flow and melt/solid interfaces in
floating-zone crystal growth under microgravity. J. Crystal Growth 102
(1990), 1043
Lan90a
D. Langbein: Crystal growth from liquid columns. J. Crystal Growth
104 (1990), 47
Lan90
Y. Kumashiro and E. Sakuma: The preparation of single crystals of
refractory carbides and nitrides. J. Crystal Growth 52 (1981), 597
Kum81b
Y. Kumashiro, A. Itoh, A. Sakuma and S. Misawa: An apparatus for RF
plasma zone melting. J. Crystal Growth 52 (1981), 495
Kum81a
R. Krishnamurti: Some further studies on the transition to turbulent
convection. J. Fluid Mech. 60 (1973), 285
Kri73
I. Kramer, D. Langbein and M. Harr: Study on possible improvements of
the performance of the automatic mirror heating facility (AMF). ESA
Contract Report, Contract No. 6065/84/NL/PR(SC) (Battelle, Frankfurt
1985)
Kra85
Zh. Kozhoukharova and S. Slavchev: Computer simulation of the
thermocapillary convection in a non-cylindrical floating zone. J. Crystal
Growth 74 (1986), 236
Koz86
5. References 233
O. A. Louchev, S. Otani and Y. Ishizawa: The incorporation of convec-
tion in 1D models of float zone and traveling solvent techniques. J.
Crystal Growth 167 (1996), 333
Lou96
O.A. Louchev: The influence of natural convection on the formation of a
molten zone under optical heating. J. Crystal Growth 133 (1993), 261
Lou93b
O.A. Louchev: Onset of supercooling in front of the solidification inter-
face in zone melting materials processing. J. Crystal Growth 131, (1993),
209
Lou93a
J. Liu and R.H. Zee: Growth of molybdenum-based alloy single crystals
using electron beam zone melting. J. Crystal Growth 163 (1996), 259
Liu96
K.H. Lie, J.S. Walker and D.N. Riahi: Melt motion in the float zone
process with an axial magnetic field. J. Crystal Growth 109 (1991), 167
Lie91
K.H. Lie, J.S. Walker and D.N. Riahi: Free surface shape and AC electric
current distribution for float zone silicon growth with a radio frequency
induction coil. J. Crystal Growth 100 (1990), 450
Lie90
J. Li, J. Sun and Z. Saghir: Buoyant and thermocapillary flow in liquid
encapsulated float zone. J. Crystal Growth 131 (1993), 83
Li93
M. Levenstam and G. Amberg: Hydrodynamical instabilities of thermo-
capillary in a half-zone. J. Fluid Mech. 297 (1995), 357
Lev95
M. Levenstam, G. Amberg, E. Tillberg and T. Carlberg: Weak flows in a
floating zone configuration as asource of radial segregation. J. Crystal
Growth 104 (1990), 641
Lev90
I.A. Lesk, A. Baghdadi, R.W. Gurtler, R.J. Ellis, J.A. Wise, M.G.
Coleman: Ribbon-to-ribbon crystal growth. In: 12th IEEE Photovoltaic
Specialists Conf. Record, New York 1976, 173
Les76
N. DeLeon, J. Guldberg and J. Salling: Growth of homogeneous high
resistivity FZ silicon crystals under magnetic field bias. J. Crystal
Growth 55 (1981), 406
Leo81
H. Lenski: Advanced facilities for crystal growth. Microgravity
Quarterly 1 (1990), 47
Len90
V.B. Lazarev: Experimental study of surface tension of indium-antimony
alloys. Russian J. Phys. Chem. 38 (1964), 172
Laz64
S. Lauer: Strstellen in Galliumantimonid. Ph.D. thesis Universitt
Freiburg i. Br., to be submitted 1997
Lau97
C.W. Lan: Thermal capillary analysis of floating-zone growth of tube
crystals: Steady-state and conduction dominated calculations. J. Crystal
Growth 135 (1994), 606
Lan94b
C.W. Lan: Heat transfer, fluid flow, and interface shapes in the floating-
zone growth of tube crystals. J. Crystal Growth 141 (1994), 265
Lan94a
234 5. References
A. Mhlbauer (Ed..): Technological data. In: Technology of Si, Ge and
SiC, Eds. M. Schulz and H. Weiss. Landolt-Brnstein New Series,
III/17c. (Springer, Berlin 1984)
Mh84
A. Mhlbauer, W. Erdmann and W. Keller: Electrodynamic convection
in silicon floating zones. J. Crystal Growth 64 (1983), 529
Mh83
T.E. Morthland and J.S. Walker: Thermocapillary convection during
floating-zone silicon growth with a uniform or non-uniform magnetic
field. J. Crystal Growth 158 (1996), 471
Mor96
T. Mizutani, K. Matsumi, H, Makino, T. Yamamoto and T. Kato: Single
crystal growing apparatus using infrared heating. NEC Res. and Dev. 33
(1974), 86
Miz74
A.C. Metaxas and R.J. Meredith: Industrial microwave heating. (Peter
Peregrinus Ltd., London 1988)
Met88
A.N. van Mau, C. Ance and G. Bougnot: Elaboration de monocristaux
dantimoniure de gallium par zone fondue flottante. Mat. Res. Bull. 3
(1968), 901
Mau68
K. Matsumoto, Y. Fujimori, M. Shimizu, R. Usami, T. Kusunose, H.
Kimura, M. Ohyama, S. Ishikura, H. Nishida, N. Negishi and S.
Kawabata: Telescience testbed examination aboard japanese experiment
module (JEM): Life and material science experiments. Acta Astronautica
27 (1992), 167
Mat92
I. Martinez and A. Crll: Liquid bridges and floating zones. In: Fluid
Physics, Lecture Notes of Summer Schools, Eds. M.G: Velarde and C.I.
Christov (Vol.5 Ser.B of World Scientific Series on Nonlinear Science
(Ed. L.O. Chua), World Scientific, Singapore 1995), 175
Mar95
I. Martinez, A. Sanz, J.M. Perales and J. Meseguer: Freezing of along
liquid column on the TEXUS-18 sounding rocket flight. ESA Journal 12
(1988), 483
Mar88
I. Martinez and A. Eyer: Liquid bridge analysis of silicon crystal growth
experiments under microgravity. J. Crystal Growth 75 (1986), 535
Mar86
W. Mackie and C.H. Hinrichs: Preparation of ZrC
x
single crystals by an
arc melting floating zone technique. J. Crystal Growth 87 (1988), 101
Mac88
Y.S. Luh, R.S. Feigelson, M.M. Fejer and R.L. Byer: Ferroelectric
domain structures in LiNbO
3
single-crystal fibers. J. Crystal Growth 78
(1986), 135
Luh86
.L. Lube: Modern methods of monitoring and control in crystal growth.
In: Growth of Crystals, Vol. 13, Ed. E.I. Givargizov (Consultants
Bureau,/Plenum Publishing Corp., New York 1986), 353
Lub86
5. References 235
H. Naumann and G. Schrder: Bauelemente der Optik. 5th Ed. (Hanser,
Mnchen 1987)
Nau87
D.O. Nason, C.T. Yen and W.A. Tiller: Measurement of optical proper-
ties of some molten oxides. J. Crystal Growth 106 (1990), 221
Nas90
S. Nakamura, K.Kakimoto and T. Hibiya: Convection visualization and
temperature fluctuation measurement in a molten silicon column. In:
Materials and Fluids under Low Gravity, Proc. of the IXth Europ. Symp.
on Gravity-Dependent Phenomena in Physical Sciences, Berlin 1995, L.
Ratke, H. Walter and B. Feuerbacher Eds. (Springer, Berlin 1996), 343.
Nak96
S. Nakamura and T. Hibiya: Thermophysical properties data on molten
semiconductors, Int. J. Thermophysics 13 (1992), 1061
Nak92
I. Nakatani, K. Masumoto, S. Takahashi, I. Nishidi, T. Kiyoawa and N.
Koguchi: Growth of single crystal InSb by floating zone method. J. Jpn.
Inst. Metals 54 (1990), 1024
Nak90
A. Murgai, H.C. Gatos and W.A. Westdorp: Effect of microscopic
growth rate on oxygen microsegregation and swirl defect distribution in
Czochralski-grown silicon. J. Electrochem. Soc. 126 (1979), 2240
Mur 79
A. Murgai, H.C. Gatos and A.F. Witt: Quantitative analysis of microseg-
regation in silicon crystal growth by the Czochralski method. J. Electro-
chem. Soc. 123 (1976), 224
Mur76
G. Mller and A. Ostrogorsky: Convection in melt growth. In: Handbook
of Crystal Growth 2b, Ed. D.T.J. Hurle (Elsevier/North-Holland, Amster-
dam 1994), 709
Ml94
G. Mller and R. Rupp: The role of Marangoni convection in the growth
of GaAs crystals by the floating zone technique under microgravity.
Crystal Properties and Preparation 35 (1991), 138
Ml91
G. Mller: Convection and inhomogeneities in crystal growth from the
melt. In: Crystals: Growth, Properties and Applications 12, Ed. H.C.
Freyhardt. (Springer, Berlin 1988)
Ml88
G. Mller, G. Neumann and H. Matz: A two-Rayleigh-number model of
buoyancy-driven convection in vertical melt growth configurations. J.
Crystal Growth 84 (1987), 36
Ml87
G. Mller and H. Jacob (Eds.): Technological data. In: Technology of
III-V, II-VI and non-tetrahedrally bounded compounds, Eds.M. Schulz
and H. Weiss. Landolt-Brnstein New Series, III/17c. (Springer, Berlin
1984)
Ml84
A. Mhlbauer, A. Muiznieks, J. Virbulis, A. Ldge and H. Riemannn:
Interface shape, heat transfer and fluid flow in the floating zone growth
of large silicon crystals with the needle-eye technique. J. Crystal Growth
151 (1995), 66
Mh95
236 5. References
J.M. Quenisset and R. Naslain: Automatisation dun appareil de fusion
de zone vertical. J. Crystal Growth 30 (1975), 169
Que75
M. Przyborowsky, T. Hibiya, M. Eguchi and I. Egry: Surface tension
measurement of molten silicon by the oscillating drop method using
electromagnetic levitation. J. Crystal Growth 151 (1995), 60
Prz95
R.P. Poplawsky and J.E. Thomas: Floating zone crystals using an arc
image furnace, Rev. Sci. Instrum. 31 (1960), 1303
Pop60
J.A.F. Plateau: Experimental and theoretical researches on the figures of
equilibrium of a liquid mass withdrawn from the action of gravity. In:
Smithsonian Institution, Annual Report 1863 (Government Printing
Office, Washington 1863), 207
Pla63
W. G. Pfann: Zone melting. 2nd Ed.(J. Wiley, New York 1966) Pfa66
W.G. Pfann and D.W. Hagelbarger: Electromagnetic suspension of a
molten zone. J. Appl. Phys. 27(1956), 12
Pfa56
W.G. Pfann: Principles of zone melting. J. Metals, Trans. AIME 194
(1952), 747
Pfa52
M. Palaniswamy, F.D. Gnanam, P. Ramasamy and G.S. Laddha: Growth
of Li
2
CO
3
single crystals by zone melting technique. Cryst. Res. Technol.
17 (1982), 911
Pal82
S. Otani, S. Honma, Y. Yajima and Y. Ishizawa: Preparation of LaB
6
single crystals from boron-rich molten zone by the floating zone method.
J. Crystal Growth 126 (1993), 466
Ota93
S. Otani and Y. Ishizawa: Preparation of LaB
6
single crystals by the
traveling solvent floating zone method. J. Crystal Growth 118 (1992),
461
Ota92
A.G. Ostrogorsky and G. Mller: Model of the effective segregation
coefficient applied to low.convective solidification in microgravity. J.
Crystal Growth 128 (1993), 207
Ost93
A.G. Ostrogorsky and G. Mller: A model of the effective segregation
coefficient, accounting for convection in the solute layer at the growth
interface. J. Crystal Growth 121 (1992), 587
Ost92
Z. Niu, K. Mukai, Y. Shiraishi, T. Hibiya and K. Kakimoto: Effects of
temperature and oxygen on the surface tension of molten silicon. In:
Proc. 4th Asian Thermophysical Properties Conf., Tokyo 1995, 73
Niu95
L. Neumann and R.A. Huggins: Technique for the zone melting of
insulators. Rev. Sci. Instrum 33 (1962), 433
Neu62
R.J. Naumann: Modelling flows and solute redistributrion resulting from
small transverse accelerations in Bridgman growth. J. Crystal Growth
142 (1994), 253
Nau94
5. References 237
F. Rosenberger: Fundamentals of crystal growth I: Macroscopic equilib-
rium and transport concepts. (Springer, Berlin 1979)
Ros79
G.D. Robertson and D.J. O'Connor: Magnetic field effects on float-zone
Si crystal growth. III: Strong axial fields. J. Crystal Growth 76 (1986),
111
Rob86b
G.D. Robertson and D.J. O'Connor: Magnetic field effects on float-zone
Si crystal growth. II: Strong transverse fields. J. Crystal Growth 76
(1986), 100
Rob86a
D. Rivas, J. Sanz and C. Vzquez: Temperature field in a cylindrical
crystal heated in a mono-ellipsoid mirror furnace. J. Crystal Growth 116
(1992), 127
Riv92
I. da Riva and I. Martinez: Floating zone stability. Proc. 3rd European
Symp. on Material Science in Space, Grenoble 1979 (ESA-SP142), 67
Riv79
H. Riemann, A. Ldge, B. Hallmann and T. Turschner: Einflu schwa-
cher bis mittlerer transversaler Magnetfelder auf das Wachstum von
Float-Zone Si-Kristallen. Jahresbericht 1995, Institut fr Kristallzch-
tung (IKZ), Berlin
Rie95a
H. Riemann, A. Ldge , K. Bttcher, H.-J. Rost, B. Hallmann, W. Schr-
der, W. Hensel and B. Schleusener: Silicon floating zone process:
Numerical modelling of RF field, heat transfer, thermal stress, and
experimental proof for 4 inch crystals. J. Electrochem. Soc. 142 (1995),
1007
Rie95
D.N. Riahi and J.S. Walker: Float zone shape and stability with the
electromagnetic body force due to a radio-frequency induction coil. J.
Crystal Growth 94 (1989), 635
Ria89
W.K. Rhim, S.K. Chung, A.J. Rulison and R.E. Spjut: Thermophysical
properties measurement of molten silicon by high temperature electro-
static levitation. In: Proc. of the 4th Asian Thermophysical Properties
Conference, Tokyo 1995, 353
Rhi95
S.F. Ray: Applied photographic optics. (Focal Press, London 1988) Ray88
Lord Rayleigh: On the capillary phenomena of jets. In: Proc. of the
Royal Soc. of London (London 1879), 71
Ray79
A. Ramos, H. Gonzalez and A. Castellanos: Liquid bridges subjected to
electric fields. In: Fluid Physics, Lecture Notes of Summer Schools, Eds.
M.G: Velarde and C.I. Christov (Vol.5 Ser.B of World Scientific Series
on Nonlinear Science (Ed. L.O. Chua), World Scientific, Singapore
1995), 193
Ram95
D.H.H. Quon, S. Chehab, J. Aota, A.K. Kuriakose, S.S.B. Wang, M.Z.
Saghir and H.L. Chen: Float zone crystal growth of bismuth germanate.
Microgravity Quarterly 3 (1993), 135
Quo93
238 5. References
H. Schildknecht: Zonenschmelzen. (Verlag Chemie, Weinheim 1964) Sch64
E. Scheil: Bemerkungen zur Schichtkristallbildung. Z. Metallkunde 34
(1942), 70
Sch42
E.M. Savitsky and G.S. Burkhanov: Growth of single crystals of high
melting metal alloys and compounds by plasma heating. J. Crystal
Growth 43 (1978), 457
Sav78
M. Saurat and A. Revcolevschi: Preparation by the floating zone method,
of refractory oxide monocrystals, in particular gallium oxide, and study
of some of their properties. Revue Internationale des Haute Temperatures
et des Refractaires 8 (1971), 291
Sau71
G.A. Satunkin, V.A. Tatarchenko and V.I. Shaitanov: Determination of
the growth angle from the shape of a crystal lateral face and solidified
separation drops. J. Crystal Growth 50 (1980), 133
Sat80
H. Sasaki, Y. Anzai, X. Huang, K.Terashima and S. Kimura: Surface
tension variation of molten silicon measured by the ring method. Jpn. J.
Appl. Phys. 34 (1995), 414
Sas95
A. Sanz, J. Meseguer and L. Mayo: The influence of gravity on the
solidification of a drop. J. Crystal Growth 82 (1987), 81
San87
A. Sanz: The crystallization of a molten sphere. J. Crystal Growth 74
(1986), 642
San86
M. Salk, M. Fiederle, K.W. Benz, A.S. Senchenkov, A.V. Egorov and
D.G. Matioukhin: CdTe and CdTe
0.9
Se
0.1
crystals grown by the travelling
heater method using a rotating magnetic field. J. Crystal Growth 138
(1994), 161
Sal94
R. Rupp and G. Mller: Experimental study of the surface tension of
molten GaAs and its temperature dependence under controlled As-vapor
pressure. J. Crystal Growth 113 (1991), 131
Rup91
R. Rupp: ber die Herstellung von GaAs mit dem tiegelfreien
Zonenschmelzverfahren unter besonderer Bercksichtigung der thermo-
kapillaren Konvektion. Ph.D. thesis, Universitt Erlangen-Nrnberg
1990
Rup90
R. Rupp, G. Mller and G. Neumann: Three-dimensional time-dependent
modelling of the Marangoni convection in zone melting configurations
for GaAs. J. Crystal Growth 97 (1989), 34
Rup89
A.M.J.G. van Run: Computation of the non-steady motion of the silicon
crystal-melt interface due to temperature fluctuations in the melt close to
this interface. J. Crystal Growth 54 (1981), 195
Run81
A.M.J.G. van Run: Computation of striated impurity distributions in
melt-grown crystals taking account of periodic remelt. J. Crystal Growth
47 (1979), 680
Run79
5. References 239
R.W. Series and D.T.J. Hurle: The use of magnetic fields in semiconduc-
tor growth. J. Crystal Growth 113 (1991), 305
Ser91
V.N. Semenov, B.B. Straumal, V.G. Glebovsky and W. Gust: Prepara-
tion of Fe-Si single crystals and bicrystals for diffusion experiments by
the electron-beam floating zone technique. J. Crystal Growth 151 (1995),
180
Sem95
H.G. Sell and W.M. Grimes: Modes of operation of an electron beam
floating zone melting furnace: Growth of single crystals of metals and
alloys. Rev. Sci. Instrum. 35 (1964), 64
Sel64
L.E. Scriven and C.V. Sternling: The Marangoni effects. Nature 187
(1960), 186
Scr60
D. Schwabe, X. Da and A. Scharmann: Unstable flow and solidification
speed due to the interaction of thermocapillary and solutocapillary forces
in directional solidification. J. Crystal Growth 166 (1996), 483
Sch96
M. Schweizer: GaAs-Kristalle gezchtet mit freien Schmelzzonen.
Masters thesis Universitt Freiburg i.Br. 1995
Sch95
R. Schwarz, A.N. Danilewsky, G. Bischopink and K.W. Benz: Crystal
growth in the zone melting facility (ZMF). Proc. VIIIth Europ. Symp. on
Materials and Fluid Sciences in Microgravity, Bruxelles 1992
(ESA-SP333, 1992), 703
Sch92
W. Schrder, H.-J. Rost and E. Wolf: Investigations for the improvement
of the radial doping homogeneity of dislocation-free floating zone silicon
crystals. Cryst. Res. Technol. 24 (1989), 3
Sch89
D. Schwabe: Surface-tension driven flow in crystal growth melts. In:
Crystals - Growth, Properties and Applications, Ed. H.C. Freyhardt.
(Springer, Berlin 1988), 75
Sch88
D. Schwabe, A. Scharmann and F. Preisser: Studies of Marangoni
convection in floating zones. Acta Astronautica 9 (1982), 183
Sch82
D. Schwabe: Marangoni effects in crystal growth melts. PhysicoChemi-
cal Hydrodynamics 2 (1981), 263
Sch81
D. Schwabe, A. Scharmann and F. Preisser: Steady and oscillatory
Marangoni convection in floating zones under 1g. In: Proc. 3rd European
Symp. on Materials Sciences in Space, Grenoble1979. (ESA-SP142),
327
Sch79b
D. Schwabe and A. Scharmann: Some evidence for the existence and
magnitude of a critical Marangoni number for the onset of oscillatory
flow in crystal growth melts. J. Crystal Growth 46 (1979), 125
Sch79a
D. Schwabe, A. Scharmann, F. Preisser and R. Oeder: Experiments on
surface tension driven flow in floating zone melting. J. Crystal Growth
43 (1978), 305
Sch78
240 5. References
T. Surek and S.R. Coriell: Shape stability in float zoning of silicon
crystals. J. Crystal Growth 37 (1977), 253
Sur77
T. Surek: Theory of shape stability in crystal growth from the melt. J.
Appl. Phys. 47 (1976), 4384
Sur76b
T. Surek: The meniscus angle in germanium crystal growth from the
melt. Scripta Metallurgica 10 (1976), 425
Sur76a
T. Surek and B. Chalmers: The direction of growth of the surface of a
crystal in contact with its melt. J. Crystal Growth 29 (1975), 1
Sur75
A.J. Strauss: The physical properties of CdTe. Rev. Phys. Appl. 12
(1977), 167
Str77
R.N. Storey and R.A. Laudise: Use of hollow cathode dc plasma
discharge float zoning for the growth of materials with high melting
points: The growth of single crystals of Ta
2
C. J. Crystal Growth 6
(1970), 261
Sto70
W.G. Smith, W.A. Tiller and J.W. Rutter: A mathematical analysis of
solute redistribution during solidification. Canad. J. Phys. 33 (1955), 723
Smi55
I. Shindo, N. II, K. Kitamura and S. Kitamura: Single crystal growth of
substituted yttrium iron garnets Y
3
Fe
5-x
(Ga,Al)
x
O
12
by the floating zone
method. J. Crystal Growth 46 (1979), 307
Shi79
X.F. Shen, A.V. Anilkumar, R.N. Grugel and T.G. Wang: Utilizing
vibration to promote microstuctural homogeneity during floating-zone
crystal growth processing. J. Crystal Growth 165 (1996), 438
She96
X.F. Shen, R.N. Grugel, A.V. Anilkumar and T.G. Wang: The influence
of controlled surface streaming on thermocapillary convection during
float-zone processing. In: Microstructural Design by Solidification
Processing, Eds. E.J. Lavernia and M.N. Gungor. (The Minerals, Metals
and Materials Society, London 1992), 173
She92
R. Shetty, R. Balasubramaniam and R. Wilcox: Surface tension and
contact angle of molten semiconductor compounds II. gallium arsenide.
J. Crystal Growth 100 (1990), 58
She90
J.A. Shercliff: Thermoelectric magnetohydrodynamics. J. Fluid. Mech.
91/2 (1979), 231
She79
J.S. Sha: Zone refining and its applications. In: Crystal Growth, Ed. B.R.
Pamplin (2nd Ed.., Pergamon Press, Oxford 1980), 301
Sha80b
J.S. Sha: Environment for crystal growth. In: Crystal Growth, Ed. B.R.
Pamplin (2nd Ed.., Pergamon Press, Oxford 1980), 105
Sha80a
Y.M. Shaskow and N.Y. Shushlebina: Effect of electromagnetic mixing
of a melt on the growth of silicon single crystals. Fiz. Khim. Obrab.
Mater. 1 (1972), 34
Sha72
5. References 241
E. Tillberg and T. Carberg: The influence of convection on the solute
distribution in crystals grown by the floating-zone technique. Appl.
Microgravity Tech. II (1989), 2
Til89
W.A. Tiller, K.A. Jackson, J.W. Rutter and B.Chalmers: The redistribu-
tion of solute atoms during the solidification of metals. Acta Met. 1
(1953), 428
Til53
T.Y. Tien and R.M. Garretson: A floating zone single crystal growing
apparatus. J. Crystal Growth 16 (1972), 177
Tie72
D. Thevenard, A. Rouzaud, J. Comra and J.J. Favier: Influence of
convective thermal oscillations on a solidification interface in Bridgman
growth. J. Crystal Growth 108 (1991), 572
The91
H.C. Theuerer: Method of processing semiconductive materials. US
patent No. 3060123, 17.12.1952
The52
A. Tegetmeier, A. Crll, A. Danilewsky and K.W. Benz: GaSb: surface
tension and floating-zone growth J. Crystal Growth 166 (1996), 651
Teg96
A. Tegetmeier, A. Crll, M. Schweizer, P. Dold and K.W. Benz: Surface
tension and growth angle: two important parameters of the floating zone
process. Proc. of the 4th Asian Thermophysical Properties Conference
(Tokyo 1995), 105
Teg95a
A. Tegetmeier: Oberflchenspannung und Wachstumswinkel: zwei
wichtige Parameter der Schmelzzonenzchtung. Ph. D. thesis Universitt
Freiburg i. Br. 1995
Teg95
A. Tegetmeier, A. Crll and K.W. Benz: A formula describing the
temperature dependence of surface tension for some semiconductor
melts. J. Crystal Growth 141 (1994), 451
Teg94
R.E. Taylor, L.R. Holland and R.K. Crouch: Thermal diffusivity
measurements on some molten semiconductors. High Temperatures -
High Pressures 17 (1985), 47
Tay85
V.A. Tatarchenko: Shaped Crystal Growth. In: Handbook of Crystal
Growth 2b, Ed. D.T.J. Hurle (Elsevier/North-Holland, Amsterdam 1994),
1011
Tat94
D.Y. Tang and R.K. Route: Growth of Barium Metaborate (BaB
2
O
4
)
single crystal fibers by the laser heated pedestal growth method. J.
Crystal Growth 91 (1988), 81
Tan88
K. Takagi and M. Ishii: Growth of LaB
6
single crystals by a laser heated
floating zone method. J. Crystal Growth 40 (1977), 1
Tak77
E.M. Swiggard: Liquid encapsulated vertical zone melt (VZM) growth of
GaAs crystals. J. Crystal Growth 94 (1989), 556
Swi89
242 5. References
J.D. Verhoeven, E.D. Gibson, M. Noack and R.J. Conzemius: An arc
floating zone technique for preparing single crystal lanthanum hexabo-
ride. J. Crystal Growth 36 (1976), 115
Ver76
R. Velten, D. Schwabe and A. Scharmann: The periodic instability of
thermocapillary convection in cylindrical liquid bridges. Phys. Fluids A3
(1991), 267
Vel91
R. Velten, D. Schwabe and A. Scharmann: Gravity-dependence of the
instability of surface-tension-driven flow in floating zones. In: Proc.
VIIth European Symp. on Materials and Fluid Sciences in Microgravity,
Oxford 1989 (ESA-SP-295), 271
Vel89
P. Vanek, S. Kadeckova, M. Jurisch and W. Lser: Melt convection
driven by electrodynamic forces from induction heating. J. Crystal
Growth 63 (1983), 191
Van83
H.P. Utech and M.C. Flemings: Elimination of solute banding in indium
antimonide crystals by growth in a magnetic field. J. Appl. Phys. 37
(1966), 2021
Ute66
H. Ueda: Resistivity striations in germanium single crystals. J. Phys. Soc.
Japan 16 (1961), 61
Ued61
C.E. Turner, N.H. Mason and A.W. Morris: The growth of barium titan-
ate crystals by the travelling solvent zone technique. J. Crystal Growth
56 (1982), 137
Tur82
B. M. Turovski and M.G. Milvidski: Besonderheiten des Kristall-
wachstums bei der Zchtung nach dem Czochralski-Verfahren aus der
Schmelze. Phys. Abhandlungen aus der Sowjetunion 6 (1962), 199.
Original paper in: TRR 3 (1961), 2519
Tur62
H. Tsuiki: Temperature distribution near the molten zone in a floating
zone imaging furnace: effect of absorptivity. J. Crystal Growth 84
(1987), 467
Tsu87
Z. Trousil: Glow discharge as a heat source for surface and floating zone
melting. Czechslov. J. Phys. B 12 (1962), 227
Tro62
D. Trivich and G.P. Pollack: Preparation of single crystals of cuprous
oxide in an arc image furnace. J. Electrochem.Soc. 117 (1970), 344
Tri70
P. Tison, D. Camel, I. Tosello and J.J. Favier: Experimental and theoreti-
cal study of Marangoni flows in liquid metallic layers. In: Proc. of the 1st
Int. Symp on Hydromechanics and Heat/Mass Transfer in Microgravity,
Perm-Moscow 1991 (Gordon and Breach, Philadelphia 1992), 121
Tis92
W.A. Tiller and C.T. Yen: Some consequences of a strong interface
field-effect operating during the growth of TiO
2
-alloy crystals from the
melt. J. Crystal Growth 109 (1991), 120
Til91
5. References 243
B. Xiong and W.R. Hu: Crystal growth in floating zone with phase
change and thermo-solutal convections. J. Crystal Growth 125 (1992),
149
Xio92
A.F. Witt, M. Lichtensteiger and H.C. Gatos: Experimental approach to
the quantitative determination of dopant segregation during crystal
growth on a microscale: Ga doped Ge. J. Electrochem. Soc. 120 (1973),
1119
Wit73
K.Th. Wilke and J. Bohm: Kristallzchtung. (VEB Deutscher Verlag der
Wissenschaften, Berlin 1988, and under license, Harri Deutsch, Thun
1988).
Wil88
L.O. Wilson: The effect of fluctuating growth rates on segregation in
crystals grown from the melt; I and II. J. Crystal Growth 48 (1980), 435
Wil80b
L.O. Wilson: Analysis of microsegregation in crystals. J. Crystal Growth
48 (1980), 363
Wil80a
L.O. Wilson: On interpreting a quantity in the Burton, Prim and Slichter
equation as a diffusion boundary layer thickness. J. Crystal Growth 44
(1978), 247
Wil78
A.A. Wheeler: The effect upon Czochralski growth of periodic modula-
tion of the growth rate. J. Crystal Growth 56 (1982), 67
Whe81
H. Wenzl, A. Fattah, D. Gustin, M. Mihelcic and W. Uelhoff: Measure-
ments of the contact angle between melt and crystal during Czochralski
growth of gallium and germanium. J. Crystal Growth 43 (1978), 607
Wen78
Measurements of the contact angle between melt and crystal during
Czochralski growth of copper. J. Crystal Growth 36 (1976), 319
Wen76
L.R. Weisberg and G.R. Gunther-Mohr: Radiant energy heater. Rev. Sci.
Instr. 26 (1956), 896
Wei56
T. Watson: Simulation der Temperaturprofile bei der Kristallzchtung
von Halbleitern in einem Monoellipsoid-Spiegelofen. Masters thesis
Universitt Freiburg i. Br. 1994
Wat94
M.J. Wargo and A.F. Witt: Numerical modelling of electric current
induced growth layers generated during Czochralski pulling. J. Crystal
Growth 66 (1984), 541
War84
Z.Q. Wang and and D. Stroud: Monte Carlo study of liquid GaAs: Bulk
and surface properties. Phys. Rev. B 42 (1990), 5353
Wan90
Z.Q. Wang and S. Stroud: Monte Carlo studies of liquid semiconductor
surfaces: Si and Ge. Phys. Rev. B42 (1988), 1384
Wan88
H.U. Walter: A mechanism for generation of pulsating growth and
nonrotational striationsduring initial transient of solidification. J. Electro-
chem. Soc. 123 (1976), 1098
Wal76
244 5. References
W. Zuhlehner: Status and future of silicon crystal growth. Mat. Sci. Eng.
B4 (1989), 1
Zuh89
G. Ziegler: Zur Bildung von versetzungsfreien Siliciumeinkristallen. Z.
Naturforsch 16a (1961), 219
Zie61
S.N. Zadumkin: Surface tension and heat of sublimation of silicon,
germanium and tin. Sov. Phys. Solid State 1 (1959), 516
Zad59
G.W. Young and A. Chait: Surface tension driven heat, mass, and
momentum transport in a two-dimensional float-zone. J. Crystal Growth
106 (1990), 445
You90
G.W. Young and A. Chait: Steady-state thermal-solutal diffusion in a
float zone. J. Crystal Growth 96 (1989), 65
You89
C.T. Yen and W.A. Tiller: Incorporating convection into
one-dimensional solute redistribution during crystal growth from the
melt: II. The initial transient solution. J. Crystal Growth 129 (1993), 224
Yen93
C.T. Yen and W.A. Tiller: Incorporating convection into
one-dimensional solute redistribution during crystal growth from the
melt: I. The steady-state solution. J. Crystal Growth 118 (1992), 259
Yen92
A. Yeckel, A.G. Salinger and J.J. Derby: Theoretical analysis and design
considerations for float-zone refinement of electronic grade silicon
sheets. J. Crystal Growth 152(1995), 51
Yec95
J. Yangyang, Z. Shuqing, H. Yijing, Z. Hangwu, L. Ming and H.
Chaoen: Growth of Lithiumtriborate (LBO) single crystal fiber by the
laser heated pedestal growth method. .J. Crystal Growth 112 (1991), 283
Yan91
B. Xiong and W.R. Hu: Influence of low gravity level on crystal growth
in flaoting zone. J. Crystal Growth 133 (1993), 155
Xio93
5. References 245
246 5. References

Potrebbero piacerti anche