Sunteți pe pagina 1din 16

Solidification and Spangle Formation of Hot-Dip-Galvanized

Zinc Coatings
J. STRUTZENBERGER and J. FADERL
Solidification of hot dip coatings was studied regarding thermal conditions. Optical phenomena occurring at the liquid zinc surface were documented and the solid zinc surface was characterized in
respect to optical and microscopic appearance, distribution of Pb and Al, crystal orientation, and
topography. Resulting from these observations, a solidification model can be derived: zinc nucleation
occurs at the steel/zinc interface. Due to thermal conditions in the slightly undercooled liquid zinc
film, solidification occurs by rapid sideways dendritic expansion of the nucleated grains along the
steel/zinc interface.
Dendritic growth is controlled by interaction of crystal orientation of the nucleated zinc grain and
thermal conditions in the undercooled layer. This leads to formation of different shaped grains with
thicker and thinner sectors. The mechanism of sideways expansion continues until the entire interface
is covered with dendritic zinc grains. Even though the zinc outer surface is still a liquid phase, final
spangle size, as well as surface appearance and shape of the grains, is already determined at that
early stage of solidification. Further growth only leads to a thickening of the solid layer; however,
its relief remains almost unchanged. Thickening occurs relatively slowly due to the fact that marginal
heat flow toward the surface now represents the limiting factor.
Growth of the solid zinc layer results in continuous enrichment of Pb and Al in the residual liquid.
Then, outer surface solidification starts as segments of single grains emerge. Distribution of the
enriched residual melt in between the already solid areas depends on the relief of the solid layer.
Finally, eutectic Zn-Pb reaction with precipitation of Pb particles takes place, which defines the dull
appearance of these regions. Solidification for lead-free coatings is essentially the same, except
that the final eutectic Zn-Pb reaction is missing. Additional investigations of dendritic secondary arm
spacing indicate that Pb does not act by suppressing zinc nucleation. Pronounced dendritic growth
is proposed to be favored by a change in interfacial energy. The new solidification model is applicable
for a wide range of processing conditions and explains the origin of the typical spangle structure.

I.

INTRODUCTION

GALVANIZED sheet is manufactured commercially by


hot dipping steel sheet in a galvanizing bath to produce a
thin zinc-rich protective coating on the steel. The structure
and the appearance of the obtained coating layer are influenced by the alloying elements added to the galvanizing
bath. Industrially used zinc baths typically contain 0.15 to
0.25 wt pct Al, and Pb up to 0.2 wt pct Al is added to
provide the unwelcome reaction of Fe and Zn by building
a thin interfacial layer consisting of the intermetallic compound Fe2Al5. Lead is known to produce a structure consisting of very large grains, termed spangles; at
concentrations greater than 0.03 to 0.04 wt pct.
Until recently, spangle morphology was the only criteria
for determining acceptability of hot-dip-galvanized steel,
which was mainly used in the construction industry. Nowadays, galvanized sheet steel is extensively used for appliance and automotive applications. Accordingly, the quality
requirements for galvanized sheet steel are becoming increasingly stringent because of the higher coating performance demanded, particularly for painted exposed panels.

Because of the more demanding uses of hot-dip-coated


products, a better understanding of the galvanizing process,
especially of the solidification and spangle formation, is
important to control spangle related technological properties such as surface reactivity, cracking behavior, or paint
adhesion.
Up to now, a lot of investigations[16] have been performed, but the formation of the spangle, its origin, and the
growth characteristics were widely speculative, and the
knowledge of zinc solidification was only fragmentary. The
aim of the present work was to supplement previous findings by new observations and to conclude a practicable solidification model for thin zinc layers, which can explain
the correlation between crystal orientation, bath chemistry,
growth characteristics, surface appearance, and spangle
size.
The present article is mainly based on a Masters thesis
by one of the authors (JS), which was performed at
VOEST-ALPINE Stahl Linz GmbH in cooperation with the
Montanuniversitat Leoben.
II.

J. STRUTZENBERGER, Postgraduate Student, and J. FADERL, Senior


Engineer, Metallic Coatings, are with the Research and Development
Department, VOEST-ALPINE STAHL LINZ GmbH, A-4031 Linz,
Austria.
Manuscript submitted February 5, 1997.
METALLURGICAL AND MATERIALS TRANSACTIONS A

EXPERIMENTAL PROCEDURE AND


RESULTS

A. Thermal Conditions
Concerning zinc solidification, the amount of undercooling occurring in the liquid layer, and the temperature graVOLUME 29A, FEBRUARY 1998631

Fig. 1Schematic illustration of the HDG simulator (for better


understanding the cooling jets and the wiping jets are drawn in a 90 deg
tilted position).

Table I.

Correlation between Cooling Rate, Solidification


Period, and Undercooling

Cooling
Rate

Atmosphere

Solidification
Period*

Undercooling*

2 K/s
4 K/s
15 K/s

air
5 pct H2/N2
5 pct H2/N2

10 s
5s
2s

,1 K
,1 K
,2 K

*Results obtained from two different bath compositions (Zn 1 0.2 wt


pct Al 1 0.0025 wt pct Pb and Zn 1 0.2 wt pct Al 1 0.05 wt pct Pb)
did not vary significantly; thus, an average value is given.

dient within the zinc film were at the center of interest.


Therefore, temperature measurements were made during
cooling and solidification of the galvanized coating after
hot dipping.
Experiments with 1-mm cold-rolled steel panels (100 3
200 mm) were performed in a hot dip galvanizing (HDG)
simulator, which was built by the Belgian steel research
institute, CRM-IRM. The construction of this apparatus is
shown in Figure 1. The water-cooled glove box is equipped
with a sample lock (up to 40 samples), an infrared annealing furnace (maximum 1000 7C), and cooling and wiping
devices. The vessel containing the zinc smelter is resistance
heated. The sample, infrared furnace, and pot cover can be
moved by pneumatic pistons. The HDG simulator carries
out galvanizing cycles that comprise a preheating, an intermediate cooling, an immersion in the zinc bath, a gas wiping, an optional reheating, and a final cooling. The
galvanizing sequence is controlled by a programmable
command module.
During a galvanizing cycle, the sample temperature, the
smelter temperature, and the gas temperature are controlled
632VOLUME 29A, FEBRUARY 1998

by thermocouples, the sample temperature is additionally


monitored by a pyrometer. The gas flow rates for cooling
and wiping and the sample speed are also controlled and
monitored. The dew point and O2 content of the protective
N2 1 H2 atmosphere within the glove box are measured at
the gas exhaust.
The samples underwent the standard HDG simulation cycle, including annealing, cooling to bath temperature, immersion, withdrawal, wiping, and controlled jet cooling
with 5 pct H2/N2 atmosphere. Additional tests were carried
out with pregalvanized 1-mm steel panels. The samples
were immersed using the pneumatic drive of the HDG simulator and held for about 20 seconds in the bath. Temperature measurements by thermocouples welded to the steel
sheet ensured that the original coating was remelted and
replaced by new bath material. After withdrawal, the samples were wiped and cooled in the air.
All samples were coated with a zinc layer of roughly 25
mm in thickness. Cooling rates at the sheet surface varied
from 2 K/s (natural convective cooling in air) up to 15 K/s
(jet cooling with a gas flow rate of 10 m3/h-5 pct H2/N2
atmosphere); the latter represents typical processing conditions on the industrial galvanizing line. Temperature
curves for cold rolled as well as for pregalvanized samples
were recorded with Ni-CrNi thermocouples, which were
welded to the surface of the steel panels. Measured undercooling and solidification periods are given in Table I.
According to the aforementioned cooling rates, the corresponding temperature gradients in the liquid zinc film
prior to solidification were approximated. Due to the symmetry of the steel/coating system, only one side was considered:
steel panel:
zinc layer:

0.5 mm
25 mm

equals
equals

3.935 kg/m2
0.179 kg/m2

Combined with the specific heat capacities of iron (621.1


J/kgzK)[7] and liquid zinc (479.9 J/kgzK),[7] at 419.5 7C, an
average heat capacity for one-half of the coated sample was
calculated:
C sample
5 2530 J/Kzm2
p
From the coated sample, heat can only be removed by
conduction trough the liquid zinc film. The total heatflow
through the zinc film depends on the average cooling rate
of the steel/coating system, which was approximated by the
cooling rate recorded at the steel surface (minimum 5 2
K/s and maximum 5 15 K/s). The heat flow is given by
Eq. [1]:
DT
z
Q 5 C sample
p
Dt

[1]

DT
z
Q 5 2lZn-liquid
Dx

[2]

(lZn-liquid 5 57.7 J/mzszK at 419.5 7C[8])


The temperature gradients in the liquid zinc film prior to
solidification are given by Eq. [2]. Considering a 25-mm
zinc coating, the approximated differences in temperature
between the steel/zinc interface and the zinc surface seem
to be rather small:
METALLURGICAL AND MATERIALS TRANSACTIONS A

Fig. 2Optical micrograph: spangle finished surface with dimpled,


ridged, orthogonal-dendritic, shiny, and feathery areas.

Fig. 4Optical micrographridged area.

Fig. 5Scanning electron microscopic surface imageorthogonaldendritic area.


Fig. 3Scanning electron microscopic surface imagedimpled area.

DTmin 5 2.2z1023 K
DTmax 5 1.6z1022 K

for a cooling rate of 2 K/s


for a cooling rate of 15 K/s

As a consequence of these slight differences in temperature and the slight undercooling, heterogeneous nucleation
at the rough steel/zinc interface, which offers preferred nucleation sites for zinc dendrites, is proposed (for homogeneous nucleation, a much greater amount of supercooling
would be necessary[9]).
B. Optical and Microscopical Surface Characterization
1. Surface structure types
The typical spangle finish of zinc coatings is well known.
Shiny, dull, and feathery areas can be recognized with the
naked eye (Figure 2). Macroscopic dull appearing areas
could be subdivided into three microscopically different
forms.
a. Dull areas
Dimpled: The dimpled structure (Figure 3) is characterized by spherical Pb particles, which are randomly spread
across the uneven zinc surface.
METALLURGICAL AND MATERIALS TRANSACTIONS A

Ridged: Precipitated Pb particles are arranged in lines


between structured zinc ridges (Figure 4).
Orthogonal-dendritic: The orthogonal-dendritic morphology (Figure 5) is characterized by secondary dendritic arms,
which lie at 90 deg to the trunk from which they emanate.
Precipitated Pb particles are distributed between the dendritic arms.
b. Shiny areas
Shiny regions invariably display a more or less pronounced hexagonal dendritic structure (Figure 6). Pb precipitations are missing almost completely.
c. Feathery areas
Feathery structures consist of a lathlike pattern with
alternate shiny dendritic branches and dull areas (Figure 7).
2. Arrangement of structure types
A typical zinc spangle is shown in Figure 8. Hexagonal
dendritic structure is clearly developed, especially within
the shiny sector. Shiny regions always coexist, as shown in
Figure 8, with dimpled/dull regions within one grain. The
shiny and dimpled regions are separated either by a straight
dendritic arm boundary or by a feathery sector lying in
between. The pronounced hexagonal dendritic structure,
VOLUME 29A, FEBRUARY 1998633

Fig. 6Scanning electron microscopic surface imageshiny area.

Fig. 8Scanning electron microscopic surface imageshiny-dull divided


spangle.

sheet, which was coated with a 0.21 wt pct Al and 0.075


wt pct Pb bath, were analyzed. Figure 9 shows the strong
enrichment of Pb and Al in the dimpled/dull appearing
regions. The average Pb concentrations for dimpled structured sectors (given in Table II) correspond to the eutectic
composition of 0.9 wt pct according to the phase diagram
of the system Zn-Pb (Figure 10).
By knowing the Pb concentration of the galvanizing bath
used, the thickness of the dimpled appearing eutectic ZnPb layer can be determined:
cbath 5 0.075 wt pct
ceutectic 5 0.9 wt pct

Fig. 7Scanning electron microscopic surface imagefeathery area.

which emanates in the shiny part of the grain, continues


underneath the dimpled/dull area, and frequently shows
through. In the dull part of the spangle, the zinc dendrite
is covered by the dimpled structure. In areas where the zinc
dendrite is not totally covered by the dimpled structures,
single shiny branches jut out and alternate with dull lines
to display the feathery appearance.
Ridged structures frequently coexist with indistinct feathery structures within one grain. The orthogonal-dendritic
morphology invariably extends over the entire grain surface.
Conclusions: Dull appearance is causally related to precipitated Pb particles, which are arranged either interdendritically or in larger closed areas for dimpled structures.
Lead-free coatings appear shiny throughout due to the
absence of Pb precipitations.
C. Distribution of Pb and Al across the Coating Surface
The distribution of Pb and Al across the coating surface
was determined by scanning electron microprobe analysis.
Each measured spot represented an area of 60 mm in diameter. The information depth was about 2 to 3 mm.
Two parallel samples of an industrial galvanized steel
634VOLUME 29A, FEBRUARY 1998

Considering the dimpled area extending over half of a


shiny-dimpled/dull divided spangle, the average thickness
of the eutectic layer is about 2 to 4 mm, which corresponds
with the dimensions of the spherical Pb precipitations. Due
to the small amount of Pb taking part in the eutectic reaction, the eutectic does not develop a regular lamellar or
rodlike form. The eutectic zinc phase crystallizes at the primary zinc dendrite, and Pb precipitates in the form of
spherical particles at the coating surface.
Conclusions: Dull appearing areas within shiny-dull divided spangles originate from the eutectic Zn-Pb reaction.
Therefore, solidification of dimpled areas occurs at the end
of the total solidification period.
D. Spangle Shape and Crystallographic Orientation
Spangle finished samples were manufactured in the
HDG-simulator using a 0.2 wt pct Al and 0.05 wt pct Pb
bath. X-ray diffraction experiments were performed with
selected spangles to determine the correlation between crystallographic orientation, surface morphology, and shape of
the grain.
Zinc solidification occurs by dendritic crystallization.
Dendrite arms grow in preferred crystallographic directions,
which are the fast growth directions, and the most commonly observed directions for zinc are ^1100& and more
rarely ^1120& and ^0001&.[2,3,4] The ^1100& and ^1120& directions lay within the (0001) basal plane of the zinc lattice.
The perfect zinc spangle displays a star-shaped denMETALLURGICAL AND MATERIALS TRANSACTIONS A

Table II.

Sample 1
Sample 2

Pb Concentrations in Dimpled Areas


Analyzed Area
in mm2

Pb Concentration
in Wt Pct

0.97
1.30

0.55 (range: 0 to 14.67)


1.67 (range: 0 to 30.53)

(a)

Fig. 10Phase diagram of the system Zn-Pb (schematic).[10]

(b)

(c)
Fig. 9Concentration mapping across a typical shiny-dimpled divided
spangle: (a) for Pb, (b) for Al, and (c) optical micrograph.

dritic morphology with 60 deg angles between the dendritic


arms, which correspond to the fast growing ^1100& direcMETALLURGICAL AND MATERIALS TRANSACTIONS A

tions. The basal plane is aligned (almost) perfectly parallel


to the sheet surface. If the basal plane is more or less misaligned with respect to the sheet plane, the surface appearance and the shape of the spangle changes.[3,4,5] To clarify
the correlation, five selected grains were analyzed with respect to the crystallographic orientation, surface appearance, and shape (Figures 11 through 15). The {0002} and
{1101} pole figures were recorded to determine the crystallographic grain orientation. The stereographic projections
correspond to the micrographs as they are oriented in the
figures. The inclination axis of the basal plane with respect
to the steel sheet is marked in the micrographs. Additionally, the elongation of the spangle, which occurs in the
direction of the inclination axis, was measured for each
grain.
The recorded pole figures confirm that each spangle represents a singlethough possibly imperfectzinc crystal;
thus, the words spangle and grain are used synonymously in the present article. A comparison of micrographs
to the corresponding pole figures shows a strong relation
between crystal orientation and spangle appearance. Elongation of the grain in the direction of the axis increases
with rising angles of inclination as long as the spangle displays the hexagonal structure. For spangles with a straight
shiny-dimpled separation line, the ^1100& dendritic arm
boundary correlates with the inclination axis and, therefore,
the direction of elongation. If the inclination axis differs
from a ^1100& dendritic direction, feathery regions are located between the shiny and the dimpled sector. For both
cases, the (0001) plane normal is invariably inclined to the
dull side of the spangle.
As the inclination of the basal plane with respect to the
steel sheet surface increases, the angle between the primary
dendritic arms gets wider; nevertheless, the hexagonal
VOLUME 29A, FEBRUARY 1998635

Fig. 11Correlation between surface appearance and crystallographic orientation for a star-shaped grain: (a) micrograph, (b) {0002} pole figure, (c)
{1101} pole figure, and (d ) schematic picture of the position of the zinc crystal with respect to the sheet plane.

Fig. 12Correlation between surface appearance and crystallographic orientation for a shiny-dimpled divided grain: (a) micrograph, (b) {0002} pole
figure, (c) {1101} pole figure, and (d ) schematic picture of the position of the zinc crystal with respect to the sheet plane.

structure is clearly developed. From inclinations of over 30


to 40 deg, the hexagonal structure becomes indistinct and
636VOLUME 29A, FEBRUARY 1998

surface appearance of the spangle changes to ridged morphology. A perpendicular position of the (0001) plane with
METALLURGICAL AND MATERIALS TRANSACTIONS A

Fig. 13Correlation between surface appearance and crystallographic orientation for a shiny-dimpled divided grain: (a) micrograph, (b) {0002} pole
figure, (c) {1101} pole figure, and (d ) schematic picture of the position of the zinc crystal with respect to the sheet plane.

Fig. 14Correlation between surface appearance and crystallographic orientation for a ridged grain: (a) micrograph, (b) {0002} pole figure, (c) {1101}
pole figure, and (d ) schematic picture of the position of the zinc crystal with respect to the sheet plane.

respect to the sheet surface results in an orthogonal-dendritic appearing grain surface.


METALLURGICAL AND MATERIALS TRANSACTIONS A

In considering the dendritic growth of a spangle, it was


pointed out that dendrites in zinc grow primarily in ^1100&
VOLUME 29A, FEBRUARY 1998637

Fig. 15Correlation between surface appearance and crystallographic orientation for an orthogonal-dendritic grain: (a) micrograph, (b) {0002} pole figure,
(c) {1101} pole figure, and (d ) schematic picture of the position of the zinc crystal with respect to the sheet plane.

basal directions. For this to be true, for spangle growth with


25-mm coating thickness and roughly 10-mm diameter, the
basal plane would have to be aligned (almost) perfectly
parallel to the steel sheet. Experimentally, great angles to
the sheet plane were found, which implies that dendritic
growth is controlled by the interaction of different crystallographically preferred growth directions. Due to that interaction, the 60 deg symmetry, which is known from basal
oriented spangles, is no longer found for grains with inclined (0001) plane, and the angle between the dendritic
arms widens.
In conclusion, the possible positions of the (0001) basal
plane with respect to the sheet plane and the resulting types
of spangles are illustrated in Figure 16. It should be remarked that the change from one spangle type to another
does not occur abruptly but continuously with increasing
inclination of the basal plane.
Conclusions: Hexagonal zinc crystallizes dendritically
with preferred growth directions, which are ^1100& and
more rarely ^1120& and ^0001&. Growth in the ^0001& direction results in an orthogonal-dendritic spangle appearance. The distribution of Pb particles within one spangle
depends on crystal orientation, especially on the inclination
of the basal plane. The rising inclination of the (0001) plane
results in increasing elongation of the grain along the inclination axis.
E. Documentation of Surface Phenomena during
Solidification
Pregalvanized steel samples were immersed into the
HDG-simulator bath and held there until the original coat638VOLUME 29A, FEBRUARY 1998

ing was remelted and replaced by new bath material. To


remove the excess zinc after withdrawal, wiping with air
was performed. Coating thickness varied between 20 and
30 mm. The galvanized steel sheet was cooled by natural
air convection resulting in cooling rates of roughly 2 K/s.
Phenomena occurring at the liquid zinc surface were observed and photographically documented.
As shown previously, cooling rates of 2 K/s gave solidification plateaus of roughly 10-second duration. Within
this period, several changes in the surface of the zinc film
were noticed. After withdrawal from the bath, the uneven
zinc film leveled to a mirrorlike shiny surface. As the solidification temperature was reached, a relief giving an anticipation of the final spangle structure appeared at the still
liquid outer surface (Figure 17). This relief remained almost
unchanged for several seconds. Then, surface solidification
started as single sectors of zinc grains emerged. These parts
of the coating surface represented the shiny areas after finished solidification. Later on, dull appearing regions solidified last, which confirms the consideration of eutectic
reaction.
Particularly, the correlation between the crystal orientation of a grain and its solidification characteristic is remarkable. The last 2 to 3 seconds of total solidification
period are illustrated in Figure 18. Area 1 represents a typical shiny-dimpled/dull divided spangle with a basal plane
slightly inclined to the steel sheet plane. Figures 18(a)
through (c) show the emergence of the finally shiny appearing half of the spangle. Even the directions of the primary dendritic branches are clearly visible. Beginning with
Figure 18(d), eutectic surface solidification with precipitation of Pb particles sets in for the rest of the grain and
METALLURGICAL AND MATERIALS TRANSACTIONS A

of a relief, which gives an anticipation of the final spangle


structure. Solidification characteristics of each spangle depend on the latters crystal orientation.
F. Surface Topography
Encouraged by the results of the optical surface observations during solidification, topography measurements of
the spangle finish were performed by Rodenstock LASER
STYLUS RM 600 system.
Figure 19 gives a comparison of optical appearance and
surface profile. It came to light that the highest laying parts
correlate with shiny appearing areas of the surface. Within
the shiny sector, the coating thickness increases continuously from the shiny-dull separation line to the grain
boundary. The dimpled/dull appearing areas represent the
lowest parts of the grain. Within the dimpled sector, the
coating thickness changes marginally.
It should be remarked that the surface profile given in
Figure 19 was obtained from an HDG-simulator sample
with a rather high coating thickness of roughly 40 mm. For
typical 10- to 20-mm coatings, the correlation between optical appearance and topography is essentially the same;
only the absolute differences in thickness are smaller.
III.

SOLIDIFICATION MODEL

A. Validity

Fig. 16Inclination of the basal plane and resulting types of spangles.

Fig. 17Relief at the liquid zinc surface during the solidification period.

defines the dull sector. Area 2 represents an orthogonaldendritic grain and area 3 a star-shaped grain. For both,
solidification occurs simultaneously across the entire spangle. Precipitated Pb is arranged between the dendritic arms
throughout the spangle surface.
Conclusions: Observations confirmed that zinc nucleation occurs at the steel/zinc interface. Advanced solidification and determination of spangle size and shape at an
early stage of solidification are indicated by the appearance
METALLURGICAL AND MATERIALS TRANSACTIONS A

The following model is based on many solidification experiments, which cover a wide range of processing conditions. Bath chemistry was varied in respect to Al and Pb
content. The lower limit of Al concentration was set at 0.16
wt pct to provide ZnFe phase formation[1,11] and the upper
limit was 0.4 wt pct. Within these limits, no essential
change in solidification sequence was observed. Concerning
the upper limit, the model is proposed to be applicable as
long as no eutectic Zn-Al reaction takes place, which means
Al concentrations of up to 1 wt pct according to the phase
diagram.[10]
The Pb content was varied between 0.0025 wt pct, which
represents so called lead-free composition, and 0.075 wt
pct. Due to marginal solid solubility, Pb is only found in
the form of precipitated particles, which characterize the
optical surface image. Galvanizing with Pb contents below
0.03 to 0.04 wt pct leads to a shiny layer finish throughout.
Coatings obtained from baths with the previous Pb contents
were found to display the typical spangle finish. Although
Pb addition markedly changes spangle size and surface appearance of the zinc coating, the solidification model is applicable within the investigated range of Pb bath
concentrations. The model explains the correlation between
Pb content and surface appearance as well as the origin of
the typical spangle structure.
Coating thickness and cooling rate were also varied in
wide ranges. Samples with coatings between 10 and 35 mm
in thickness showed no essential difference in solidification
process. Cooling rates varied from 2 K/s (natural convective air cooling) to 15 K/s (jet cooling). Within this range
also, no essential change in solidification characteristic was
noticed.
Strip entry temperature changed slightly (bath temperature 510 7C), but for all experiments, matching of strip
VOLUME 29A, FEBRUARY 1998639

Fig. 19Surface profile of a typical shiny-dimpled divided spangle.

Fig. 18(a) through ( f ) Photographically documented sequence of


surface solidification (dark areas represent the liquid zones, and light areas
the solid ones).

temperature and bath temperature before withdrawal was


presupposed. The influence of a higher difference between
bath temperature and strip entry temperature and the effect
of an application of zinc dust or water droplets to the liquid
zinc film are subjected to further investigations.
B. Solidification Process
In a typical continuous galvanizing process, the steel
sheet is cleaned (alkaline degreasing, electrolytic pickling,
etc.), annealed, cooled to bath temperature, and immersed.
After withdrawal from the zinc bath and removal of excess
zinc by wiping, the strip enters the cooling zone, where
solidification of the coating layer occurs.
Due to slight undercooling prior to the onset of solidification and the marginal difference in temperature within
the liquid film, heterogeneous zinc nucleation takes place
at the interface steel/zinc, which offers preferred nucleation
sites. This happens even though it is slightly cooler on the
surface of the liquid zinc film. The nucleated zinc crystals,
which are randomly oriented with respect to the sheet plane,
start to grow dendritically within the slightly undercooled
layer. Solidification is accompanied by the release of latent
melting heat, which must be removed by slow conduction
along the flat temperature gradient toward the layer surface.
This provides further cooling of the liquid zinc layer above
the already solid zinc crystal. Further growth toward the
surface is therefore hindered.
In areas, where the steel is not covered with flat dendritic
640VOLUME 29A, FEBRUARY 1998

crystals, unhindered heat flow toward the surface is still


possible and the undercooling in these areas increases. This
thermal constellation leads to a rapid sideways expansion
of nucleated zinc crystals along the steel/zinc interface (Figure 21(a)). The mechanism of sideways expansion continues until the entire interface is covered with dendritic zinc
grains, and therefore, further lateral heat flow is impossible.
Cameron and Harvey[3] found flat zinc dendrites attached to
the steel substrate by removing the liquid zinc immediately
after solidification had started, which confirms the accuracy
of the present considerations.
Concerning the final shape and appearance of the growing zinc grain, its crystal orientation with respect to the
sheet plane plays an important roll at the first stage of solidification. Zinc solidification occurs by dendritic crystallization and the fast, and therefore preferred, growth
directions are the six ^1100&, which lay within the basal
plane. If a fast growing ^1100& direction is aligned (almost)
perfectly parallel to the steel/zinc interface, it coincides
with the direction of maximum heat flow. This leads to fast
lateral dendritic expansion within the first stage of solidification. For grains with basal orientation, all six ^1100&
directions lay parallel to the sheet plane, which means equal
thermal growth conditions for all of them. The basal orientation therefore results in a symmetrical star-shaped grain
with 60 deg angles between the six dendritic arms, which
mark the ^1100& directions of the zinc lattice.
For crystals with inclined basal plane, thermal conditions
for the ^1100& directions differ within one grain. Near the
inclination axis, the fast growing ^11 00& directions fit well
with the directions of maximum heat flow parallel to the
steel/zinc interface. Rapid dendritic growth takes place and
causes the typical elongation of the grain along the inclination axis. At both sides of the axis, the ^1100& directions
deviate from the sheet plane. Due to thermal conditions,
growth is only allowed parallel to the interface. Directions
other than the dominant ^1100& ones must therefore additionally be involved in the sideways expansion process.
This leads to a decrease in growth velocity at both sides of
the inclination axis and causes the loss of 60 deg symmetry
of the dendritic arms. The observed elongation of grains
with inclined basal plane can be well explained by the proMETALLURGICAL AND MATERIALS TRANSACTIONS A

Fig. 20Growth conditions for grains with inclined basal plane: (a)
perpendicular to the inclination axis and (b) parallel to the inclination axis.

posed interaction of thermal conditions and crystal orientation.


Not only the size and shape but also the profile of the
dendritic grain are determined within the first stage of solidification. Again, the inclination of the basal plane is determinant. For crystals with inclined basal plane, thickness
of the solid dendritic layer differs within one grain after the
first stage of solidification. For one-half of a grain with
inclined basal plane (illustrated in Figure 20(a)left of the
inclination axis), the ^11 00& directions, which are involved
in lateral growth, deviate from the plane of maximal heat
flow, showing a component toward the surface. Due to the
strong preference of the ^1100& directions concerning
growth velocity, sideways expansion is accompanied with
a thickening of the solid layer. Within the grain, thickness
increases continuously from the inclination axis to the grain
boundary. At the other half of the grain (shown in Figure
20(a)right of the inclination axis), the ^1100& directions,
which could promote thickening, are positioned almost opposite to the direction of maximum heat flow. Thickening
is therefore hindered.
Figure 20(b) illustrates growth conditions parallel to the
inclination axis. Fast growing ^1100& directions correlate
with thermal caused growth directions, and rapid sideways
expansion can occur.
For grains with their basal plane positioned perpendicular
with respect to the sheet, dendritic expansion normal to the
inclination axis is sustained by ^0001& directions. Branching
parallel to the axis occurs in faster growing ^1100& or
^1120& directions, leading to an elongation of the grain
along the axis. After sideways expansion is finished, thickness of the solid dendritic layer is roughly constant across
the grain.
Grains with basal orientation behave similarly. All six
^1100& directions are aligned parallel to the sheet, which
results in a star-shaped dendritic crystal with roughly constant thickness across the entire grain.
The ongoing solidification process along the steel/zinc
interface is accompanied with a phenomenon, which beMETALLURGICAL AND MATERIALS TRANSACTIONS A

Fig. 21Schematic illustration of the solidification sequence for lead


containing zinc layers: (a) rapid lateral expansion, (b) slow thickening
process, (c) emergence of shiny sectors, and (d ) eutectic Zn-Pb
solidification.

comes visible at the still liquid outer surface. A relief giving


an anticipation of the final spangle structure emerges immediately after solidification starts. This phenomenon may
be explained by a gravity induced flow of melt on the vertically positioned sample during the lateral stage of solidification. This flow seems to stagnate rather quickly as the
liquid film above the solidifying grains becomes thinner
due to the thickening process of the solid zinc layer beneath. The surface tension of the melt seems to be insufficient to level the relief of the liquid outer surface during
the thickening process.
All processes mentioned previously take place within
some tenths of a second. Final size, shape, and appearance
of the growing spangle, nevertheless, are widely determined
in these early stages of solidification. The first solidification
stage ends, as the entire steel/zinc interface is covered by
a solid layer consisting of profiled dendritic grains. Further
growth is controlled by the heat flow toward the outer surface, which is limited by the removal of heat from the
coated sample by convection and thermal radiation. This
limiting mechanism leads to a rather slow thickening of the
solid zinc grain layer (Figure 21(b)). During this thickening
process, the relief of each single grain remains essentially
the same. However, the growth rate toward the surface differs for different grains with different crystallographic oriVOLUME 29A, FEBRUARY 1998641

Fig. 22Arrangement of Pb particles: (a) shiny-dimpled divided grain,


and (b) star-shaped or orthogonal-dendritic grain.

Fig. 23Spangle size vs Al and Pb: bath temperature, 470 7C; immersion
temperature, 470 7C 5 10 7C; immersion time, 6 s; cooling rate, 4 K/s;
coating thickness, 20 mm.

entations. Due to the limited heat flow, this second stage


of the solidification process takes the greater part of the
total solidification time.
Surface solidification starts when the areas that developed the greatest thickness during the sideways expansion
and the thickening process reach the outer surface of the
liquid zinc film (Figure 21(c)). Considering a grain with
inclined basal plane, the dendritic sector with ^1100& directions having a component toward the surface represents the
area, which is located closest to the liquid surface after the
initial stage of solidification. Due to the thickening process,
this sector emerges first and appears shiny after solidification is complete.
Thickening of the solid zinc grain layer is accompanied
by continuous enrichment of Al and especially Pb within
the residual liquid. As the shiny sector of a grain with inclined basal plane emerges at the liquid surface, the other
half of the grain is still covered by residual melt, which
finally reaches eutectic Zn-Pb composition. Eutectic reaction takes place. The zinc phase crystallizes at the primary
zinc dendrite and lead precipitates in the form of particles
at the coating surface (Figure 21(d)). The so-obtained dimpled structure defines the dull appearing sector within a
typical shiny-dull divided spangle.
The dimpled and the shiny appearing sectors are divided
642VOLUME 29A, FEBRUARY 1998

either by a straight dendritic arm boundarythis is the case


if the inclination axis correlates with a ^1100& direction
or by a feathery area laying in between. Feathery structures
develop in areas, where the amount of eutectic melt is insufficient to totally cover the primary zinc dendrite. Infilling
occurs between the dendritic arms to produce dull appearing eutectic lines alternating with shiny dendritic branches
(Figure 22(a)).
For grains with their (0001) basal plane positioned either
parallel or perpendicular with respect to the sheet, thickness
of the dendritic layer after the initial stage of solidification
is roughly constant across the grain. Due to the homogeneous thickening, surface solidification occurs simultaneously across the entire spangle. Residual eutectic melt is
arranged between the dendritic arms all across the spangle
surface (Figure 22(b)). Star-shaped and orthogonal-dendritic structured grains therefore display a more or less dull
appearance.
After the eutectic reaction is complete, the typical spangle structure is fully developed. Only a phase transformation of the Pb precipitation, which is insignificant
concerning spangle appearance, takes place at 318.2 7C.
Coatings obtained from galvanizing baths with Pb concentrations below 0.03 to 0.04 wt pct show a slightly different characteristic within the final stage of solidification.
Due to the reduced Pb content, the amount of enriched eutectic melt is insufficient to form larger closed areas within
the spangles. Precipitated Pb particles are only found between dendritic arms, which causes a rather uniform appearance of the coating surface.
For lead-free coatings, the eutectic Zn-Pb reaction is
missing. Pb particles are only occasionally found at boundaries between dendritic zinc grains. Due to the absence of
Pb precipitations, lead-free coatings appear shiny throughout. The reduced grain size of lead-free coatings additionally contributes to the homogeneous coating finish.
C. Effect of Bath Chemistry on Spangle Size
The dependence of grain size on the amount of Pb and
Al added to the galvanizing bath is shown in Figure 23
(data sets from Table IV). The absolute values of the spangle size are in good correlation with the data reported by
Kim and Patil.[12] As shown in Figure 23, there is an interaction of Pb and Al concerning determination of the spangle size.
Spangle formation occurs when 0.03 to 0.04 wt pct Pb
is added to the galvanizing bath. It has been proposed by
several authors[3,5] that Pb suppresses zinc nucleation, leading to large diameter spangle growth reflecting the fewer
nuclei present. Spittle and Brown[6] proposed a theory in
which Pb is said to reduce the number of nuclei and increase the undercooling prior to solidification, resulting in
a coarse spangle structure.
These theories can neither give a mechanism for suppressing the zinc nucleation nor explain the influence of
other alloying elements such as Al on grain size.
Within the present study, the influence of bath chemistry
on the amount of undercooling prior to the onset of solidification was monitored. Temperature measurements with
thermocouples welded to the steel substrates were performed for HDG-simulator samples. Cooling rate after
withdrawal from the bath was 4 K/s, giving solidification
METALLURGICAL AND MATERIALS TRANSACTIONS A

Table III.

Segregation Coefficients and Surface Tension of


Alloying Elements[2]

Alloy

Segregation
Coefficient
k

Surface Tension of
Alloying Element
in N/m

Spangles

Zn-Pb
Zn-Sb
Zn-Sn
Zn-Al
Zn

,0.01
,0.01
,0.01
0.2

0.470
0.395
0.566
0.860
0.760

yes
yes
no
no
no

plateaus of roughly 5 seconds for 20-mm coatings. The Al


bath concentration was varied between 0.2 and 0.4 wt pct
and the Pb content between 0.0025 and 0.05 wt pct. Within
this range, no significant change in undercooling could be
detected. In all tests, the undercooling was below 1 7C.
These results correlate with the findings of Fasoyinu and
Weinberg,[2] who reported undercoolings smaller than 0.5
7C for pure zinc as well as for Zn 1 0.08 wt pct Pb 1 0.35
wt pct Al coatings.
With the slight undercooling found in both studies, it is
unlikely that spangle formation is directly related to large
melt undercooling caused by Pb addition, as proposed.[6]
Spangle formation is caused by the growth of large primary dendrites. In pure metals, dendritic growth is associated with thermal undercooling. In alloys, either thermal or
constitutional undercooling, or both, can lead to dendritic
growth. Constitutional undercooling results from solute
segregation at the solid/liquid interface, the amount of segregation depending primarily on the segregation coefficient
k. The alloying elements Pb and Sb, which produce large
spangles, have very low segregation coefficients (,0.01)
and corresponding high solute concentrations in the melt at
the interface. Fasoyinu and Weinberg[2] reported that ZnSn, which does not produce large spangles, also has a very
low value of k (,0.01). Therefore, they concluded that segregation by itself cannot account for large spangles.
Fasoyinu and Weinberg[2] investigated the dependence of
grain size of coatings, obtained from baths with alloy concentrations of 0.2 wt pct, on the surface tension of the liquid/vacuum interface of pure alloying elements. The
elements that produce large spangles, Pb and Sb, have
lower surface tension than Sn and Al, which do not produce
spangles (Table III).
It should be remarked that in the considered
publications,[2,13] the surface tension of the solid/liquid interface is treated as equivalent to the surface tension of the
liquid/vacuum interface. Whether this is rigorously true
could not be proved within the present study.
Alloying elements with low segregation coefficient as
well as low surface tension were found to promote the
growth of large spangles. Assuming that Pb lowers the surface tension of the solid/liquid interface, it seems rather
unlikely that Pb can act by suppressing zinc nucleation, as
proposed. If there is an influence of Pb content on nucleation at all, it should probably promote nucleation.
It is proposed that due to the solute segregated ahead of
the growing dendrite, the growth velocity of the zinc dendrite becomes higher, resulting in larger grains. This seems
to be the dominant effect concerning spangle size, if one
considers that Pb should promote zinc nucleation. For further discussion, the influence of bath chemistry on nucleaMETALLURGICAL AND MATERIALS TRANSACTIONS A

tion was assumed to be negligible, and only the correlation


between growth velocity and surface tension of the
solid/liquid interface was investigated in detail.
If it is assumed that the surface tension of the liquid/solid
interface changes with the alloy element addition in proportion to the relative mass of the two constituents, the
change would be very small for a 0.05 wt pct Pb addition.
However, the high concentration of solute ahead of the
growing dendrite could have a much larger effect. Dendritic
growth is accompanied with an extension of the solid/liquid
interface by branching, which is controlled by interfacial
energy gsl. A decrease in interfacial energy promotes
branching and reduces the dendritic tip radius, resulting in
larger growth velocity and, therefore, larger spangles.
The growth velocity of advancing dendrites in a pure
undercooled liquid has been examined by Nash and Glicksman.[14] The growth rate in steady-state dendritic growth for
small undercoolings is given by Eq. [3].
R 5 0.128 z

aliquid zDSzDH
gslzcliquid

! ~
z

DTzcliquid
DH

2,65

[3]

where
R growth velocity;
DT undercooling;
DS entropy of fusion per unit volume;
DH latent heat of fusion;
aliquid thermal diffusivity of the liquid; and
cliquid specific heat of the liquid.
This function, although it concerns pure liquids, also
seems practicable for typical zinc coatings because of the
fact that constitutional undercooling caused by solute segregation does not account for fast dendritic growth. The
only effect of solute segregation is probably the change in
interfacial energy.
The main problem with calculating the growth velocity
using Eq. [3] is the exact determination of the solid/liquid
interfacial energy gsl.
(1) The concentration of solute atoms ahead of the
solid/liquid interface, which determines interfacial energy gsl, is not constant during the initial stage of solidification. Concentration pileup is built up due to
advancing dendritic expansion and interacts with dendritic growth velocity.
(2) Interfacial energy will not consequently change with
the alloy element concentration in proportion to the relative mass of the two constituents. Alloying elements
that lower surface tension and interfacial energy are
expected to segregate at interfaces and surfaces.
(3) Typical zinc coatings contain additions of more than
one element, and the alloying elements might interact
concerning the determination of interfacial energy and
growth velocity, as assumed previously (cf. Figure 23).
Because of these reasons, a theoretical determination of
the interfacial energy gsl was left out of consideration. Instead, determining the interfacial conditions during the solidification process by characterization of the dendritic
structure of the solidified coating surface was attempted. As
shown previously, dendritic curvature and arm spacings are
influenced by surface tension gsl of the solid/liquid interface. As a representative parameter, the secondary arm
spacing l2 of orthogonal structured dendrites was chosen.
VOLUME 29A, FEBRUARY 1998643

Table IV.

Secondary Arm Spacing of Orthogonal Structured Dendrites for Different Al and Pb Concentrations in the
Galvanizing Bath

Bath Content
(Wt Pct)

Dendritic Secondary Arm Spacing l2


(mm)
AK*
IF*
Average

Spangle Size
(mm)

Al

Pb

0.196
0.195
0.415
0.410
0.309

0.0026
0.0475
0.0023
0.0498
0.0275

AK
311
1815
158
1062
161

5
5
5
5
5

IF
46
421
30
211
16

438
2650
254
2153
276

5
5
5
5
5

72
669
27
693
37

40.7
28.7
44.8
31.5
45.8

5
5
5
5
5

2.5
3.6
7.2
3.0
4.2

41.3
29.2
49.1
28.5
45.1

5
5
5
5
5

5.0
4.2
6.3
5.3
6.6

41.0
29.0
47.3
30.1
45.4

5
5
5
5
5

3.8
3.8
6.8
4.4
5.5

*At least seven measurements per bath composition: bath temperature, 470 7C; immersion temperature, 470 5 10 7C; immersion time, 6 s; cooling
rate, 4 K/s; coating thickness, 20 mm.

Fig. 25Correlation between dendritic secondary arm spacing l2 and


spangle size d for AK and IF steel substrates.

Fig. 24Scanning electron microscopic surface image of the steel


substrates: (a) AK and (b) IF.

Coatings obtained from five different baths, which varied


in respect to Al and Pb content, were investigated, and the
secondary dendritic arm spacing was measured by simple
linear image analysis. Results for two different substrates,
one Al killedlow carbon (AK) steel grade and an interstitial free (IF) grade, are listed in Table IV.
No significant difference in secondary arm spacing could
be found for the two substrates, which indicates equal interfacial conditions for both during the solidification process. Only the bath chemistry is considered to influence the
value of l2, if the other process parameters are kept constant. Therefore, an average value for each bath composi644VOLUME 29A, FEBRUARY 1998

tion was determined. Differing spangle size mainly seems


to be a result of different nucleation rates for the two grades
caused by the surface topographies of the substrates (Figure
24). Within the investigated range of galvanizing parameters and bath compositions (Al . 0.195 wt pct), the difference in reactivity to the galvanizing bath seems to be a
minor factor concerning nucleation and final spangle size.
No ZnFe crystals, which possibly may influence the nucleation, were found in cross sections through the fully solidified layer. This result was confirmed by wet chemical
analysis. As former experiments[15] with AK and IF steel
grades showed, differences can be found in the formation
of the Fe2Al5 intermetallic layer due to the steel chemistry
but no significant differences in spangle size if the surface
topographies of AK and IF are comparable.
Figure 25 clarifies the correlation between spangle size
d and dendritic arm spacing l2. Empirical functions from
obtained data, which describe the correlation between l2
and d, were calculated by linear regression.
The aim of the l2 measurements was to obtain information about gsl during solidification process. Several theories dealing with the correlation between l2 and gsl have
been published.[9,13,16] According to Kurz and Fischer,
METALLURGICAL AND MATERIALS TRANSACTIONS A

l2 5 5.5 z (Mztf )1/3


M5

gslzD1
ln (ce /co)
z
DS mz(1 2 k)z(ce2co)

tf local solidification time;


Dl diffusivity of solute in the liquid;
DS entropy of fusion per unit volume;
m slope of liquidus;
co alloy concentration;
ce eutectic concentration; and
k segregation coefficient.

d5Az

[5]

Dzt
~DSzDT
!

1/3

For zinc solidification, it is assumed that C is roughly


constant; thus, l2 can be approximated using the following
data (D ' 1027 m2/s,[17,18] tf 5 5 s, DS 5 1.05 z 106 J/m3 z
K,[7] and DT 5 1 7C). The predicted value of l2 corresponds
well to the experimentally measured data listed in Table V.
Reducing Eq. [3] by the assumption that the parameters
thermal diffusivity of the liquid aliquid, entropy of fusion DS,
latent heat of fusion DH, specific heat of the liquid cliquid,
and undercooling DT are approximately constant within the
range of investigated bath compositions leads to the following relation:
R ; 1/gsl

[6]

Combining Eqs. [5] and [6], the correlation between secondary dendritic arm spacing l2 and growth velocity R is
given as follows:
METALLURGICAL AND MATERIALS TRANSACTIONS A

~!
R
N

1/(11D)

[8]

where

As mentioned previously, constitutional undercooling


seems not to play a significant role in zinc solidification.
The second term in the formula describing M, which represents the influence of the concentration field, can therefore be replaced by 1/DT to consider the more important
thermal undercooling. Thermal undercooling DT was found
to be below 1 7C for all experiments and, therefore, was
assumed to be approximately constant.
Following Eq. [4], the final value of l2 is largely determined by the contact time between the dendritic branches
and the liquid. Within this period, known as local solidification time tf, a ripening process causes the secondary arms
to change with time into coarser, more widely spaced ones.
For directional solidification, tf is mainly determined by the
growth velocity and the temperature gradient ahead of the
advancing dendrite. But in case of zinc, solidification occurs by lateral expansion of dendrites attached to the
steel/zinc interface. This constellation allows contact between the secondary dendritic arms and the melt until solidification of the layer is finished. The definition of the
local solidification time tf following the theory for directional solidification, therefore, did not seem useful. The total solidification period was instead considered as an
approximate value for the contact time. The solidification
period did not vary significantly for the five investigated
bath compositions; thus, it was taken as constant (5 seconds) for the following considerations.

l2 5 Czgsl1/3

[7]

Relation [7] can be transformed to Eq. [9] by using Eq.


[8], which results from considerations in terms of units and
dimensions and defines the dependence of spangle size on
growth velocity and nucleation rate for lateral growth.

where

C 5 5.5 z

l2 ; R21/3

[4]

A constant;
R growth velocity;
N nucleation rate;
d spangle size; and
D dimension (D 5 2 for lateral growth).

l2 ; d21

[9]
21

The factor of proportion between R and d depends on


the nucleation rate N, which is assumed to be rather independent of the bath composition within the investigated
range but influenced by the surface topography of the substrate. Thus, the l2 vs d graphs for AK and IF grades, illustrated in Figure 25, differ markedly, whereas the
exponent m of the empirical functions describing the graphs
is nearly the same for both substrates.
The concluding Relation [9], which results from theoretical considerations, corresponds qualitatively with the empirically found functions. But due to the deviation of the
coefficient (experimentally found values of 20.2 in comparison to 21), it does not give a real satisfactory description of the experimental findings.
The deviation of the exponent m might result from special growth conditions for zinc dendrites in the thin coating
layer. The theoretical considerations mentioned previously
are mainly based on theories describing directional solidification parallel to the heat flux direction. In the case of
zinc, solidification occurs more or less two dimensionally,
and additional heat flux perpendicular to the dendritic
growth direction toward the coating surface is possible.
For the proposed mechanism of two-dimensional solidification, a more comprehensive model, including the influence of contact between the growing dendrites and the
substrate on thermal conditions and solute segregation,
seems necessary to explain the found data in detail.
IV.

CONCLUSIONS

1. Thermal undercooling and temperature gradient within


the galvanized layer are small prior to solidification.
Accordingly, zinc solidification starts from heterogeneous nuclei at the steel/zinc interface.
2. No significant influence of bath chemistry (Al additions
between 0.2 and 0.4 wt pct and Pb additions up to 0.05
wt pct) on undercooling prior to solidification could be
found.
3. Within the initial stage of solidification, maximum heat
flow within the slightly undercooled layer occurs parallel to the steel/zinc interface, which forces the nucleated zinc grains to expand sideways along the steel/zinc
interface. Sideways expansion endures until the grains
impinge upon one another, determining the final spanVOLUME 29A, FEBRUARY 1998645

4.

5.

6.

7.

8.

9.

10.

11.

gle size. The coating surface stays liquid within this


initial stage of solidification.
Final relief and shape of the spangles are determined
within the stage of rapid sideways expansion. Dendritic
growth is controlled by an interaction of thermal conditions in the undercooled layer and crystal orientation
of the nucleated zinc grain. Inclination of the basal
plane to the sheet and the deviation of the fast growing
^1100& directions from the plane of maximum heat flow
are the determining parameters. Rising inclination of
the basal plane is associated with an increasing elongation of the grain along the inclination axis, as long
as the spangle displays hexagonal structure.
Further growth toward the coating surface leads to a
thickening of the solid zinc grain layer. The slow thickening process is controlled by heat flow toward the
surface, which is limited by the removal of heat by
external cooling. Thickening of the solid zinc layer results in an enrichment of Pb, Al, and other alloying
elements in the residual melt.
The relief of the zinc grains, which develop within the
first stage of solidification, has a determining influence
on the outer morphology of the fully solidified coating.
As the solid zinc grain layer reaches the surface, enriched residual melt is arranged interdendritically for
grains with the basal plane positioned parallel to the
sheet as well as for grains having a basal plane tilted
at angles of more than approximately 40 deg to the
sheet plane. For grains with a slightly inclined basal
planewhich means inclination angles below 40
degthe residual melt forms larger closed areas.
Solidification of the enriched residual melt occurs by
eutectic Zn-Pb reaction. The Pb precipitates at the coating surface and the distribution of Pb particles across
the surface define the optical appearance of the spangle
finish.
Solidification of lead-free zinc coatings is essentially
the same, only the final eutectic Zn-Pb reaction is missing. Due to the absence of Pb precipitations, the surface
of coatings obtained from lead-free galvanizing baths
appears shiny throughout.
The alloy addition of Pb to the galvanizing bath at
concentrations greater than 0.03 to 0.04 wt pct produces spangles of large diameter in the galvanized
layer. Rising the Al content of the zinc bath from 0.2
to 0.4 wt pct results in a slight decrease of spangle size.
There is also an interaction of Pb and Al concerning
spangle size.
The Pb is proposed to lower surface energy at the
solid/liquid interface; thus, it seems rather unlikely that
Pb can act by suppressing zinc nucleation. Decreasing
interfacial energy might lead to an increase in dendrite
growth velocity, resulting in large spangles. Promoting
rapid dendritic growth seems to be the dominant effect
of Pb addition.
Grain size correlates with dendritic secondary arm
spacing, which confirms that interfacial energy during
solidification process is a determining parameter for
grain size.

646VOLUME 29A, FEBRUARY 1998

12. Theoretical considerations concerning the relation between spangle size and dendritic secondary arm spacing correlate with the experimental findings only in a
qualitative way. At present, a more detailed theoretical
explanation of the experimentally determined l2 vs d
graphs does not seem useful, because of the many assumptions necessary due to the special growth conditions of the zinc dendrites in the coating layer within
the initial stage of solidification.
ACKNOWLEDGMENTS
The authors would like to thank Professor F. Jeglitsch,
at the Montanuniversitat Leoben, for the helpful discussion.
The general assistance of R. Pree and the analytical evaluation of the samples supervised by J. Angeli, both at
VOEST-ALPINE Stahl Linz GmbH, are also gratefully
appreciated.
REFERENCES
1. R. Kiusalaas, G. Engberg, H. Klang, E. Schedin, and L. Schon:
GALVATECH 89, Proc. Int. Conf. on Zinc and Zinc Alloy Coated
Steel Sheet, 1989, The Iron and Steel Institute of Japan, Tokyo, 1989,
pp. 485-92.
2. F.A. Fasoyinu and F. Weinberg: Metall. Trans. B, 1990, vol. 21B, pp.
549-58.
3. D.I. Cameron and C.J. Harvey: Proc. 8th Int. Hot Dip Galvanizing
Conf., London, 1967, Zinc Development Association, London, 1967,
pp. 86-97.
4. Y. Fukui, M. Koda, and Y. Hirose: Tetsu-to-Hagane, 1991, vol. 77,
pp. 939-46.
5. N.J. Wall, J.A. Spittle, and R.D. Jones: Proc. 1st Int. Conf. on Zinc
Coated Steel Sheet, Munich, June 1985, Zinc Development
Association, London, 1985, pp. C1-C6.
6. J.A. Spittle and S.G.R. Brown: INTERGALVA 91 Proc. 3rd Int. Zinc
Coated Sheet Conf., Barcelona, 1991, European General Galvanizers
Association, London, 1991, pp. S4K/1-S4K/9.
7. Handbook of Chemistry and Physics, 62nd ed., CRC Press Inc., Boca
Raton, FL, 1981-82.
8. S. Kavesh: in Principles of Fabrication, Rapid Solidification
TechnologySource Book, R.L. Ashbrook, ed., ASM, Metals Park,
OH, 1983, p. 83.
9. M.C. Flemings: Solidification Processing, McGraw-Hill Book Co.,
New York, NY, 1974, pp. 146-54 and pp. 290-94.
10. Metallographic Atlas of Zinc and Zinc Alloys, Centre de Recherches
Metallurgique, Liege, Belgium, 1981, p. 2.7.24.
11. Nai-Yong Tang and G.R. Adams: Proc., The Physical Metallurgy of
Zinc Coated Steel, TMS Annual Meeting, San Francisco, CA, A.R.
Marder, ed., TMS, Warrendale, PA, 1994, pp. 41-54.
12. Yong-Wu Kim and R.S. Patil: Proc. 1st Int. Conf. on Zinc Coated
Steel Sheet, Munich, June 1985, Zinc Development Association,
London, 1985, pp. D1-D5.
13. W. Kurz and D.J. Fischer: Fundamentals of Solidification, 3rd ed.,
Trans Tech Publications Ltd., Aedermannsdorf, Switzerland, 1989,
pp. 63-89.
14. G.E. Nash and M.E. Glicksman: Acta Metall., 1974, vol. 22, p. 1291.
15. J. Faderl, W. Maschek, and J. Strutzenberger: Proc. of GALVATECH
95, Chicago, IL, 1995, Iron and Steel Society Inc., Warrendale, PA,
1995, pp. 675-85.
16. J.D. Verhoeven: Fundamentals of Physical Metallurgy, John Wiley &
Sons, New York, NY, 1975.
17. W.J. Moore and D.O. Hummel: Physikalische Chemie, 3rd ed., Walter
de Gruyter & Co, Berlin, 1983, p. 191.
18. Gmelins Handbuch der Anorganischen Chemie, System Nummer 32,
Zink Erganzungsband, 8th ed., Verlag Chemie GmbH., Weinheim/
Bergstrabe, 1956, pp. 347-48.

METALLURGICAL AND MATERIALS TRANSACTIONS A

S-ar putea să vă placă și