Sie sind auf Seite 1von 9

International Journal of Heat and Mass Transfer 127 (2018) 394–402

Contents lists available at ScienceDirect

International Journal of Heat and Mass Transfer


journal homepage: www.elsevier.com/locate/ijhmt

Effects of aluminum concentration on the formation of inhibition layer


during hot-dip galvanizing
Ting Min, Yimin Gao ⇑, Xiaoyu Huang, Zhanpeng Gong, Kemin Li, Shengqiang Ma
State Key Laboratory for Mechanical Behavior of Materials, School of Materials Science and Engineering, Xi’an Jiaotong University, 28 Xianning West Road, Xi’an,
Shaanxi Province 710049, PR China

a r t i c l e i n f o a b s t r a c t

Article history: The formation of inhibition layer (IL) during hot-dip galvanizing with Al concentration varying from 0.0
Received 11 April 2018 to 0.8 wt pct. is studied by both experiments and numerical simulations. Using EDS and XRD, the IL is
Received in revised form 19 July 2018 identified mainly as Zn-bearing Fe2Al5. SEM results reveal that the IL thickness increases with Al concen-
Accepted 6 August 2018
tration in zinc bath, and the particles of Fe2Al5 on the surface of IL formed in 0.8 wt pct. Al bath are smal-
ler than those formed in 0.2 wt pct. Al bath. On the surface of IL generated in 0.8 wt pct. Al bath there are
some tiny particles with size about tens of nanometers randomly dispersing on the larger ones, indicating
Keywords:
that the growth of IL is so fast that the growth mode changes and nucleation occurs at the IL/zinc inter-
Crystal growth
Inhibition layer
face. Using a mesoscopic model based on the lattice Boltzmann method, numerical simulations are also
Mesoscopic simulation performed to study the reactive transport phenomena during IL formation under different Al concentra-
Reactive transport tions. The simulations reveal complex coupled mechanisms between Fe and Al diffusion, Fe dissolution,
Hot-dip galvanizing as well as nucleation and growth of Fe2Al5 and the results agree with the experiments.
Ó 2018 Elsevier Ltd. All rights reserved.

1. Introduction The concentration of Al in zinc bath is about 0.13 wt pct. for


galvannealing process and about 0.2 wt pct. for galvanizing pro-
During hot-dip galvanizing process, an extremely thin layer of cess, so most of the experiments in the literature focused on
Fe-Al compound with size about tens to hundreds of nanometers, the formation and morphologies of IL in bath with Al concentra-
called inhibition layer (IL), forms between the steel substrate and tion varying from 0.1 to 0.3 wt pct. at about 460 °C [1–4,11–15].
the zinc overlay. This layer serves as a barrier to retard or inhibit The typical morphology of IL formed during galvanizing is a
the formation of brittle Fe-Zn compounds between steel and mol- double-layer structure, with the lower layer next to substrate
ten zinc [1]. Therefore, IL has a critical influence on the microstruc- containing continuous, compact, roughly equiaxed grains with
ture and properties of zinc coating. When Al concentration in the size about tens of nanometers, and the upper layer adjacent to
zinc bath is lower than 0.14 wt pct., no Fe-Al compound forms liquid zinc comprising coarser, larger, elongated grains with size
and reaction between Fe and Zn occurs, while when Al concentra- about hundreds of nanometers [1,2,4,6,11–14,18]. To develop
tion is higher than this threshold, the Fe-Al compound forms IL and more environment friendly galvanizing method, galvanizing steel
hinders the formation of Fe-Zn compound [2–4]. The IL formation with scale reduced by heating hydrogen or carbon was suggested
occurs quickly in only a few seconds, which is affected by various by researchers to replace conventional acid pickling [19–22]. The
factors including chemical constituent, temperature of molten zinc, surface of steel with reduced scale is rough and porous [23,24],
substrate material and so on [1,4–10]. Moreover, multiple sub- and a higher Al concentration (0.7 wt pct.) in zinc bath has been
processes are involved in the IL formation process, including disso- proved to be necessary to form a complete IL on such special
lution of Fe, diffusion of Fe and Al in liquid zinc and solid IL, reac- substrate surfaces [19]. However, to the best of the authors’
tion between Fe and Al as well as competition with Fe-Zn reaction knowledge, there are few articles focusing on the IL formation
[1–5,11]. Thus, study of IL formation is a challenging topic for in Zn bath containing Al concentration higher than 0.3 wt pct.
experimental and theoretical researchers and has drawn great in the past two decades [25]. Besides, as a vital process during
attention [1–18]. galvanizing, the formation of IL is affected by Al concentration
significantly. It is therefore theoretically and industrially mean-
ingful to explore the influence of Al concentration on the IL
⇑ Corresponding author.
formation.
E-mail address: ymgao@mail.xjtu.edu.cn (Y. Gao).

https://doi.org/10.1016/j.ijheatmasstransfer.2018.08.016
0017-9310/Ó 2018 Elsevier Ltd. All rights reserved.
T. Min et al. / International Journal of Heat and Mass Transfer 127 (2018) 394–402 395

As a complementary method to experiments, numerical simula- 2. Experiments


tion has been developed not only to predict galvanizing processes
at bath scale to capture the fluid flow and transport processes [26– 2.1. Experimental method
31], but also at atomic scale to explore the intrinsic properties of
compounds forming in zinc bath [32–38]. In the bath-scale simula- The specimens used in this study were polycrystalline pure
tions, emphasis was put on the evolutions of velocity, temperature iron. All of them were cut into 70 mm  10 mm  4 mm and
and concentration in the whole bath, while the complex reactive degreased in a 2.0 wt pct. alkaline solution at 80 °C for 15 min,
transport processes occurring at the surface of steel were simply which were rinsed first and then pickled in 15 vol pct. hydrochloric
set as source terms or boundary conditions [26–30]. In atomic- acid for 3 min to remove oxide and rinsed. Next the specimens
scale calculation using first principle or molecular dynamics, some were prefluxed in a solution of NH4Cl and ZnCl2 for 3 min at
useful physical parameters, such as crystal structure [36], heat 60 °C, and then were dried in heater immediately before immer-
capacity [35,37], diffusivity [38] and formation energy [32] were sion in zinc bath. The solution of NH4Cl and ZnCl2 is a traditional
obtained. However, the size of the atomic-scale simulations is lim- activating agent in galvanizing, which can clean the surface further
ited to a few nanometers. Thus, there is a gap between bath-scale and maintain an un-oxidized surface before immersion in zinc
and atomic-scale simulations. A mesoscale model is a bridge which bath. Pure zinc ingots were melted in a graphite crucible first
can connect the galvanizing process parameters, variable surface and then pure Al wire and pure iron was added after zinc ingots
state and the intrinsic properties of involved compounds together. totally melted. Four zinc baths saturated with Fe and Al were pre-
Recently, Min et al. [39] developed a mesoscale model based on pared with Al concentration as 0.0, 0.1, 0.2, 0.5, 0.8 wt pct., respec-
lattice Boltzmann method (LBM) [40–42] to numerically investi- tively, and for each bath three specimens were immersed for about
gate the reactive transport processes during IL formation down 10 s at 460 °C one by one after Al wire was completely dissolved.
to nanoscale. This model not only can take into account the The cross sections of galvanized samples were cut from the
multiple physicochemical processes, but also can capture the middle of galvanized specimens, grinded and polished for observa-
dynamic evolutions of IL structure, which is difficult to realize tion using SEM and EDS. A mixture of glacial acetic acid and hydro-
using current experimental techniques [39]. In this paper, this gen peroxide with volume ratio 4:1 was used to expose the top
model was also used to explore the effects of Al concentration on view of IL by stripping off the zinc layer. The etching solution
the IL formation. can remove zinc and leave the Fe-Al intermetallic phases [15].
The formation of IL can be divided into three steps: (1) dissolu- The proceeding solution was also used to extract the particles of
tion of Fe from substrate into zinc bath, (2) rapid growth of IL by IL. X-ray diffraction analysis was performed on top view after the
nucleation and growth of Fe-Al until surface is completely covered Zn layer was removed using Bruker D8AA25X with Cu Ka radiation
and (3) diffusion-controlled growth of IL [1,5–10]. A widely used and the monochromatic operated at 40 kV and 40 mA. The diffrac-
model of IL formation was proposed by Tang [5] in which he tion scans were performed using a parallel beam geometry X-ray
applied classic nucleation theory considering the influence of tem- diffractometer with an incident angle of 1.0°. The speed was set
perature of bath and strip, thickness of strip, immersion time, heat as 30 s/degree and the step as 0.02°.
transfer as well as fluid flow on IL formation process. The IL forma-
tion was divided into two stages, high rate of IL formation by con- 2.2. Results
tinuous nucleation of Fe2Al5 and the growth of the Fe2Al5 limited
by the supply of Al through liquid diffusion [5]. This model was Fig. 1(a) and (b) shows the top-view morphology and EDS spec-
elaborated by Giorgi [8] combining the dissolution of Fe and trum of IL after stripping the overlying zinc obtained in a 0.20 wt
growth kinetics of Fe2Al5. In this model the surface of substrate pct. Al bath at 460 °C. The typical double-layer structure of IL [4]
is covered by the nucleation and lateral growth of Fe2Al5. Several was observed from top-view (Fig. 1(a)), where some larger grains
physical parameters were artificially adjusted because of the lack disperse on the top of compact and fine ones. The EDS result fur-
of experimental and simulated values. Until now, the three stages ther proves that this layer is rich in Fe, Al and Zn (Fig. 1(b)).
of IL formation are widely accepted but different mechanisms The phase identification was implemented using GIXRD from
about each stage are continuously proposed by researchers. Some top-view after stripping the overlying zinc. As shown in Fig. 1(c),
researchers pointed out the second stage was completed by nucle- except the substrate iron, only Fe2Al5 was detected, indicating that
ation [5,7] and others believe that the substrate surface was cov- the dominant phase in IL should be Fe2Al5. Combing with EDS
ered by initial nuclei and their fast growth [8,10]. Some results, IL mainly consists of Zn-bearing Fe2Al5 compound. Fig. 1
experiments indicate that the initial nucleation is FeAl3 which then (d) displays the extracted particles of IL formed in 0.2 wt pct. Al
transforms to Fe2Al5 [4]. Different factors controlling the IL growth bath, most of which are irregular polygon particles and the size
in the third stage have also been proposed in different models varies in the range of tens to hundreds of nanometers.
including diffusion of Al in the liquid zinc in Refs. [5,7,10] and solid The SEM BES images and elemental distributions of Al in cross-
diffusion of Fe in IL in Refs. [4,6,8]. In our previous model, the sur- section of ILs formed in Zn baths containing 0.0, 0.1, 0.2, 0.5 and
face is treated as being covered by nucleation and fast growth of 0.8 wt pct. Al are shown in Fig. 2. Al-rich area is hardly examined
these nuclei; the third stage is controlled by Al diffusion in liquid when the Al concentration is 0.0 or 0.1 wt pct., and thus no ele-
zinc first and then by Fe diffusion in IL [39]. mental distribution is given for these two cases. When the Al con-
In the present study, the IL formation in liquid zinc is explored centration is 0.2 wt pct. or higher, there is a band of dark contrast
by both experiment and mesoscale simulations. The emphasis is between substrate and coating zinc because of the formation of IL.
put on the effects of Al concentration. The rest of the paper is This dark band is confirmed to be an Al-rich layer as shown in the
arranged as follows. In Section 2, experiment methodology for IL EDS mapping (Fig. 2(d, f, h)). According to the Fe-Al-Zn ternary
formation is introduced, and then the IL is analyzed using SEM, phase diagram [3], Fe2Al5 can form when the Al concentration is
EDS and XRD. In Section 3, mesoscopic numerical simulations are higher than about 0.14 wt pct. The results obtained here are in
performed, and time evolutions of IL structure, concentration fields coincidence with the ternary phase diagram [3]. Moreover, as
and IL thickness are displayed. In Section 4, the experimental and shown in Fig. 2(c, e, g), when the Al concentration increases the
simulated results are combined with each other and discussed. Al-rich layer becomes thicker. Limited by the resolution of BES in
Finally, a main conclusion is drawn in Section 5. SEM, it is hard to measure the IL thickness quantitatively.
396 T. Min et al. / International Journal of Heat and Mass Transfer 127 (2018) 394–402

500 nm
keV
(a) (b)

(c) (d)

Fig. 1. Top view morphology of inhibition layer (IL) (a) SEM, (b) EDS spectrum operated at 5 kV accelerate voltage, (c) XRD pattern of IL using GIXRD, (d) morphologies of
extracted particles in IL formed in 0.20 wt pct. Al bath at 460 °C.

To compare the morphologies of ILs formed in zinc bath with a mesoscopic model based on the LBM was developed to study IL
high and low concentrations of Al, the top-view SEM SE images formation [39]. In this model, Fe and Al mass transport in liquid
of ILs formed in 0.2 and 0.8 wt pct. Al baths are displayed in zinc as well as in solid IL were simulated using the LBM, and struc-
Fig. 3. The particles in IL forming in 0.8 wt pct. Al bath are smaller ture evolution of IL due to nucleation and crystal growth of IL was
those formed in 0.2 wt pct. Al bath, and both of ILs contain compact captured by a cellular automaton (CA) method. Different mecha-
fine polyhedral grains with some dispersing larger ones. Another nisms in the literature were explored using the model. IL Double-
difference between surfaces of these two ILs is that there are some layer structures as well as typical IL thickness were successfully
tiny particles, about tens of nanometers as marked using red1 cir- captured by the model [39].
cles, distributing in IL formed in 0.8 wt pct. Al bath. In the present study, the LB mesoscopic model is further
employed to investigate the effects of Al concentration. The com-
putational domain is a rectangle with size of 1000  1000 nm as
3. Numerical study shown in Fig. 4. Part of the domain near the left is Fe with size of
100  1000 nm, while the remaining part of the domain is liquid
3.1. Simulation method zinc. According to the phase identification by GIXRD and EDS, IL
consists mainly of Zn-bearing Fe2Al5. Therefore, following
Mesoscopic simulations are performed to simulate the coupled [5,7,10], the precipitation reaction of Fe2Al5 on the steel surface
reactive transport process during IL growth process, with emphasis in contact with the liquid zinc is described by the following
on the effects of Al concentration. After the steel sheet enters the formula
Zn-Al bath, Fe dissolves from the steel surface into the liquid zinc
1
bath. Then the precipitation reaction of Fe2Al5 takes place, which 2Fe þ 5Al ! Fe2 Al5 ; DG ¼ 283; 470 þ 84:8T J mol ð1Þ
consumes Fe and Al. Nucleation and subsequent crystal growth
of Fe2Al5 occur on the steel surface, and IL is generated [5–10]. IL where DG is the free energy change associated with this reaction,
formation is characterized by short-time, nanoscale, and multiple which is a function of temperature, T. This reaction takes place only
physicochemical processes involved, making it one of the most if two conditions are satisfied based on the Zn-Al-Fe phase diagram
challenging problems in material science. In our previous study, [3]. First, there is a threshold value of Al concentration in the zinc
bath, below which Fe2Al5 cannot be generated and the inhibition
1
For interpretation of color in Fig. 3, the reader is referred to the web version of breaks down, leading to the generation of Fe-Zn intermetallic com-
this article. pounds [3]. This threshold value is set as 0.14 wt pct. at T = 733 K
T. Min et al. / International Journal of Heat and Mass Transfer 127 (2018) 394–402 397

(a) 0.0 wt. pct. Al, SEM (b) 0.1 wt. pct. Al, SEM

(c) 0.2 wt. pct. Al, SEM (d) 0.2 wt. pct. Al, EDS

(e) 0.5 wt. pct. Al, SEM (f) 0.5 wt. pct. Al, EDS

(g) 0.8 wt. pct. Al, SEM (h) 0.8 wt. pct. Al, EDS

Fig. 2. Cross sectional morphologies and elemental distributions of ILs obtained in (a) 0.0 wt pct., (b) 0.1 wt pct., (c, d) 0.2 wt pct., (e, f) 0.5 wt pct. and (g, h) 0.8 wt pct. Al
baths.
398 T. Min et al. / International Journal of Heat and Mass Transfer 127 (2018) 394–402

8
>
> if ðC Fe Þ2 ðC Al Þ5 < ðC Fe;sat Þ2 ðC Al;sat Þ5
>
>0
>
< or C Al < C Al;critical
SFe ¼  
>
> if ðC Fe Þ2 ðC Al Þ5 P ðC Fe;sat Þ2 ðC Al;sat Þ5
>
> 2k  ðC Fe Þ2 ðC Al Þ5
>
: p 1 2
ðC Fe;sat Þ ðC Al;sat Þ5
and C Al P C Al;critical
ð2aÞ

8
>
> if ðC Fe Þ2 ðC Al Þ5 < ðC Fe;sat Þ2 ðC Al;sat Þ5
>
> 0
>
< or C Al < C Al;critical
SAl ¼  
>
> if ðC Fe Þ2 ðC Al Þ5 P ðC Fe;sat Þ2 ðC Al;sat Þ5
500 nm >
> 5k  ðC Fe Þ2 ðC Al Þ5
>
: p 1 ðC Fe;sat Þ2 ðC Al;sat Þ5
and C Al P C Al;critical

(a) 0.5 wt. pct. Al ð2bÞ

where kp is the precipitation reaction rate constant, C the molar


concentration in the bath, Csat the saturation molar concentration
and CAl,critical the threshold concentration of Al below which Fe2Al5
cannot form. Note that in the present study, C denotes the concen-
tration with unit of mol m3, while w is weight percent with unit of
wt pct. In the generated solid IL, heterogeneous diffusion containing
both lattice diffusion and grain boundary diffusion is considered.
The lattice diffusion is the slow diffusion within a grain. The grain
boundary diffusion (or short circuit diffusion) refers to diffusion
along the grain boundaries in a polycrystalline, and is greatly faster
than the lattice diffusion [4,43]. In a polycrystalline with smaller
500 nm crystal size, there are abundant of grain boundaries, leading to
greatly larger diffusivity in nano-crystals than the corresponding
bulk counterpart [13]. This is the case in the Fe–Al compound IL
(b) 0.8 wt. pct. Al
with crystal size of tens to hundreds of nanometers. Therefore, Chen
Fig. 3. Top-view morphologies of ILs after stripping zinc layer in 0.2 wt pct. (a) and et al. [4] adopted a grain boundary diffusivity 500 times higher than
0.8 wt pct. (b) Al bath at 460 °C. lattice diffusivity to theoretically estimate the IL growth. Following
our previous study [39], two values of solid diffusivity (4.0  1012
m2 s1 and 4.0  1011 m2 s1) are randomly assigned to the IL solid
nodes with 50% probability of each value.
Under a certain temperature T, the relationship between Fe sat-
urated weight percent and that of Al is as follows [10]
 
33066
ðwFe;sat Þ2 ðwAl;sat Þ5 ¼ exp 28:1  ð3Þ
T

where wi,sat is saturated weight percent of i. For more details of the


model, one can refer to our previous study [39,44].
Initially, the bath is saturated with aluminum and iron. At the
top and bottom boundaries, periodic boundary conditions are
adopted. At the right boundary, concentrations of Fe and Al are
set as the saturation concentrations [8]. At the steel interface,
boundary condition for Al is no flux boundary condition. For Fe
at this interface, if the interface is not covered by IL, the dissolution
reaction follows [8,10],

@C Fe
DFe ¼ kd ðC 0Fe;sat  C Fe Þ ð4Þ
@n

where DFe is the diffusivity of Fe in liquid zinc, n the direction nor-


Fig. 4. Evolutions of IL structures and Fe concentration for the case with Al mal to the reactive surface pointing to the void space, kd with unit of
concentration as 0.2 wt pct. m s1 the dissolution reaction rate constant, C 0Fe;sat the saturation
concentration of Fe in liquid zinc bath without Al. On the other
hand, if a computational node at interface is occupied by IL, the con-
centration of Fe there is set as the C 0Fe;sat [5,8,10]
[3,5]. Second, product of Fe and Al concentrations, namely, should
be higher than ðC Fe;sat Þ2 ðC Al;sat Þ5 , otherwise only liquid phase can
C Fe ¼ C 0Fe;sat ð5Þ
exist [3]. In other words, the crystal growth requires supersatura-
tion. Therefore, source terms for the Fe and Al concentrations SFe The values of physicochemical variables adopted in the simula-
and SAl are as follows tion are listed in Table 1.
T. Min et al. / International Journal of Heat and Mass Transfer 127 (2018) 394–402 399

Table 1
Values of physicochemical variables used in the simulations.

Physicochemical variables Values


Density of liquid Zn, qZn 6570 kg m3
Saturated weight percent Fe in liquid Zn at 0.04 wt pct.
T = 733 K, w0Fe;sat
Dissolution rate constant of Fe at the 1.75  102 m s1
Fe-liquid Zn interface, kd
Density of Fe2Al5 with dissolved Zn, qFe2 Al5 4720 kg m3
Molar volume of the IL component, VIL 6.61  105 m3 mol1
Precipitation rate constant of IL, kp 10000 mol m3 s1

3.2. Evolution of IL under various Al concentrations

Fig. 4 shows the evolutions of IL structures and Fe concentration


for the case with Al weight percent as 0.2 wt pct. In Fig. 4, black is
steel, brown triangle denotes IL, and the contour is for Fe weight
percent. Before the formation of IL, the source for the Fe into the
bath is steel dissolution given by Eq. (4), and it is expected that
weight percent of Fe decreases from bath-steel interface to the Fig. 6. Time evolutions of the average concentration of Fe in molten zinc with
right boundary (t = 0.001 s). As the precipitation reaction contin- different Al concentration.
ues, the first layer of IL forms (t = 0.2 s), and the entire steel surface
is covered by IL, leading to the supply mechanisms of Fe into the
bath shifted to the slow solid diffusion. Such slow diffusion Fig. 6 gives the time evolution of average Fe concentration in 0.2
becomes the constraint of IL growth later. Due to the heteroge- and 0.8 wt pct. Al bath. The Fe concentration increases dramati-
neous diffusivity (including lattice diffusion and grain boundary cally fast initially due to the dissolution of substrate, and then
diffusion) of Fe in the IL, IL grows relatively quicker at some sites decreases slowly to their saturation concentration resulting from
with high values of diffusivity, and the surface of IL is no longer the formation of IL, which can reduce the dissolution of Fe by cov-
smooth but becomes rough (t = 5 s). The Fe weight percent gradi- ering substrate surface and consume Fe dissolved in liquid zinc
ent is still obvious at t = 5.0 s, which means Fe still can diffuse simultaneously. At final stage the Fe concentration in 0.8 wt pct.
away from the IL/liquid zinc interface as be consumed by reaction. Al bath is slightly lower than its saturation concentration due to
Fig. 5 shows the evolutions of IL structures and the Fe weight reaction overshoot. Comparing the two cases, the dissolution time
percent in 0.8 wt pct. Al bath. It can be seen that in a very short of Fe in 0.8 wt pct. Al bath is much shorter than that in 0.2 wt pct.
time of 0.001 s the first layer of IL forms. At t = 0.1 s, more than five Al and the Fe concentration approaches to the saturation concen-
layers of IL have been generated, with effects of the heterogeneous tration earlier in 0.8 wt pct. Al bath.
diffusion clearly observed. The IL generated in the 0.8 wt pct. Al Fig. 7 further shows the solid structures of IL for four cases with
bath is much thicker than that in the 0.2 wt pct. Al bath and the different values of wAl;sat . First, as wAl;sat increases, IL grows faster
concentration gradient of Fe in zinc bath is negligible in the 0.8 and becomes thicker, coincident with the experimental results
wt pct. Al bath when immersion time is 5 s, which means all of shown in Fig. 2. Under a low wAl;sat , the upper layer of the IL is rel-
Fe diffusing from substrate is completely consumed by the IL for- atively flat; under a high wAl;sat , the upper layer of IL are coarser
mation and Fe seldom can diffuse away from the interface. agree with Fig. 2(g). Fig. 8 illustrates the time evolutions of the IL
thickness. For the case with 0.2 wt pct. Al, the thickness of IL at
t = 5 s is about 50 nm, while for the 0.8 wt pct. case, the thickness
of IL at t = 5 s can be as high as 390 nm.

4. Discussion

The experimental observation of IL formed in zinc bath results


from underlying coupled physicochemical processes, and is influ-
enced by multiple factors such as Al concentration, immersion
time, temperature, etc. Numerical simulations can provide details
of the distributions and evolutions of important variables such as
concentration, and thus serve as a complementary tool for experi-
ments. Combining the experimental observations and simulation
results, there are three typical variations of IL with Al concentra-
tion: (1) the IL grows faster as Al concentration increases, (2) the
final IL thickness of IL increases with the increasing of Al concen-
tration and (3) the particles of Fe2Al5 become smaller and some
finer particles disperse on the top morphology of IL when the Al
concentration increases to 0.8 wt pct. The IL forming in 0.2 wt
pct. Al bath has been observed and discussed by many researchers
[2,4,11–15,17,45,46] and recently it was successfully captured by
Fig. 5. Evolutions of IL structures and Fe concentration for the case with Al mesoscopic simulations [39]. Therefore, the emphasis will be put
concentration as 0.8 wt pct. on the morphology variation of IL with Al concentration between
400 T. Min et al. / International Journal of Heat and Mass Transfer 127 (2018) 394–402

Fig. 7. Evolutions of IL structure under different Al concentration.

0.2 and 0.8 wt pct. The formation of IL can be considered as three


steps: the dissolution of Fe from substrate before IL formation,
the fast growth controlled by initial nucleation and the fast growth
of the nuclei before the substrate surface is completely covered by
IL, and the slow IL growth controlled by diffusion. We will discuss
the IL formation process step by step.
The dissolution of Fe from the substrate can be described by Eq.
(4). Increasing the Fe concentration gradient at the interface can
enhance the dissolution rate. According to Eq. (3), the higher the
wAl, sat, the lower the wFe, sat. With temperature as 733 K, if wAl, sat
is 0.2 wt pct. (CAl, sat is 486.67 mol m3), wFe, sat is 0.011 wt pct.
(CFe, sat is 12.87 mol m3). As wAl, sat increases to 0.8 wt pct.
(CAl, sat is 1216.67 mol m3), wFe, sat greatly reduces to 3.4  104
wt pct. (CFe, sat is 0.40 mol m3). Because C 0Fe;sat in Eq. (4) (or
w0Fe;sat ) is a constant for a given temperature, it is obvious that a
lower wFe, sat leads to a higher Fe dissolution rate. Therefore, for
the higher wAl, sat case, Fe dissolution is accelerated. On the other
hand, the Al concentration also affects the Fe dissolution by accel-
erating nucleation and growth of Fe2Al5, which can cover the sub-
strate surface, reduce the dissolution area, and consume Fe
Fig. 8. Time evolution of the IL thickness under different Al concentration. near the interface in liquid zinc. According to Figs. 4 and 5, the
T. Min et al. / International Journal of Heat and Mass Transfer 127 (2018) 394–402 401

dissolution time before the formation of the first layer of IL varies gradually becomes insignificant, as shown in Fig. 8, where the gra-
from 0.2 s to 0.001 s, hence the amount of Fe dissolution into Zn dient of all of growth curves is close to an extremely small value.
bath reduces rapidly with Al concentration, as shown in Fig. 6. It
can be deduced from Fig. 6 that part of Fe dissolved from substrate
is precipitated in the form of Fe2Al5 and the remaining part of Fe
5. Conclusion
dissolved into zinc bath raising the Fe concentration near the inter-
face, as shown in Figs. 4 and 5.
During hot-dip galvanizing, a small amount of Al is added into
The nucleation of Fe2Al5 begins when the Fe concentration near
the molten zinc bath, leading to the formation of Zn-bearing Fe2Al5
the interface is high enough and the product of Fe and Al concen-
inhibition layer (IL) between the substrate and the zinc coating.
trations exceeds their saturated concentration product according
Investigating effects of Al concentration on the IL formation is
to Eq. (3). Here, the model improved by Dutta [9] is used to
important for both scientific researches and industrial applications.
describe the nucleation rate N. In Dutta’s model [9] the activation
In this study, both experiments and numerical simulations are con-
energy of Fe diffusion across the substrate/zinc interface is
ducted to investigate crystal growth during IL formation in bath
included. The nucleation rate N can be calculated by [5,9]
with Al concentration varying from 0 to 0.8 wt pct. Several tech-
  niques including SEM, EDS and XRD are employed to investigate
8p Dcpl
3
DG  þDGD
N ¼ n0 v 0 s  C Al exp  ; DG ¼ ð6Þ the morphologies and composition of the ILs. There is no IL formed
kT 3 DG2 in the 0 and 0.1 wt pct. Al baths, consistent with the conclusion in
the literature that below a threshold value of 0.14 wt pct. Al the IL
where n0 is the number of Fe atoms on the surface of substrate, v0
cannot form. Above this threshold value, the experimental results
the lattice vibration frequency (1013), s⁄ the total number of atoms
show that IL thickness increases as Al concentration raises. Besides,
surrounding the critical nucleus, DG⁄ the energy barrier of nucle-
it is also observed from the SEM images that the particles in IL
ation, GD the activation energy for diffusion across the substrate/
formed in 0.8 wt pct. Al bath are smaller than those formed in
melt interface (21 kJ mol1), k Boltzmann constant, T the temper-
0.2 wt pct. Al bath, while some tiny crystals distribute on the sur-
ature, and Dcpl the interface energy change for a semi-sphere
face of IL formed in the 0.8 wt pct. Al bath, indicating that the
nucleus on the substrate. According to Eq. (6), the nucleation rate
nucleation and growth occur simultaneously.
is determined by CAl and T. For a given temperature T, the nucle-
Currently, it is challenging to experimentally observe the time
ation rate N is exclusively affected by CAl. A higher Al concentration
evolution of IL, which completes in several seconds and involves
results in faster nucleation, and further leads to smaller particles of
different physicochemical processes. Simulations help to get a
Fe2Al5, which is consistent with experimental observation shown in
comprehensive understanding of evolutions and distributions of
Fig. 3. Besides nucleation of Fe2Al5, the fast growth of existing nuclei
important variables. Therefore, numerical simulations are imple-
occurs simultaneously in the second stage. The growth rate of
mented using a mesoscopic model based on the LBM. In this model,
Fe2Al5 particles vgrowth can be described by the following formula
coupled processes of Fe and Al diffusion, Fe dissolution reaction, IL
" # nucleation and crystal growth are considered. Time evolutions of
ðC Fe Þ2 ðC Al Þ5
v growth  kp 1 ð7Þ the IL structures as well as Fe concentration are presented and dis-
ðC Fe;sat Þ2 ðC Al;sat Þ5 cussed in detail. It is found that as Al concentration increases, iron
dissolution and IL growth are accelerated, leading to quicker IL for-
According to Eq. (3), ðC Fe;sat Þ2 ðC Al;sat Þ5 is constant for a given tem-
mation and thicker IL. For the physicochemical variables adopted
perature, and thus the growth rate of particles, vgrowth, raises expo-
in the present study, the final IL thickness is about 50 nm for 0.2
nentially with the Al and Fe concentrations at the interface. Due to
wt pct. Al case, which is relatively high as about 390 nm for the
higher Al concentration and the faster dissolution of Fe for the case
0.8 wt pct. case. Accordingly, the formation and structures of IL is
with higher Al concentration the growth of nuclei is accelerated
depended on Al concentration and immersion time based on the
significantly in the second stage, which results in a much shorter
Fe and Al diffusion.
time for complete coverage of substrate surface in the 0.8 wt pct.
Al case. This process completed in less than 0.001 s is very chal-
lenging, if possible, to be directly observed by experiments, but is
consistent with the top-view morphologies shown in Fig. 3. Fig. 3 Conflict of interest
shows the morphology of IL changes when Al concentration varies
from 0.2 to 0.8 wt pct. The dispersed tiny crystals sticking to the We declare that we have no financial and personal relationships
surface of IL formed in 0.8 wt pct. Al bath indicate that the concen- with other people or organizations that can inappropriately influ-
tration product of Fe and Al at the interface is very high resulting in ence our work, there is no professional or other personal interest
a faster growth mode and nucleation on the growing particles of any nature or kind in any product, service and/or company that
occurs during their growth. This phenomenon is seldom reported could be construed as influencing the position presented in, or the
in the literature [1–4,11–18,25,46,47]. review of, the manuscript entitled.
Once the surface of substrate is completely covered by Fe2Al5,
the third stage of IL formation starts. At the beginning of the third
stage, Fe near the interface is still supersaturated with Fe2Al5 as Acknowledgement
shown in Figs. 4(b) and 5(a). As the growth proceeds the Fe concen-
tration near the interface is diminished to its saturation point and Yimin Gao thanks the support of the Science and Technology
the only Fe source for reaction becomes the diffusion of Fe across IL Project of Guangdong Province in China (2015B010122003,
from substrate, as shown in Figs. 4(d) and 5(d). The IL growth is 2015B090926009), the Science and Technology Project of Guangz-
limited by the slow solid diffusion of Fe in IL. For a given thickness hou City in China (201604046009). Shengqiang Ma thanks the
of IL, the amount of Fe diffusing across IL is the same, but the National Natural Science Foundation of China (51771143). Ting
growth rate of IL in 0.8 wt pct. Al bath is larger due to the higher Min thanks Dr. Jiuhong Wang at State Key Laboratory for Manufac-
Al concentration, which is the mainly reason why IL thickness turing Systems Engineering for the help of GIXRD and Mr. Zijun
increases with Al concentration, as shown in Fig. 2. With the thick- Ren at Instrument Analysis Center of Xi’an Jiaotong University for
ening of the IL, the effect of Al concentration on the growth rate the help of SEM observation.
402 T. Min et al. / International Journal of Heat and Mass Transfer 127 (2018) 394–402

References [24] D.J. Ding, H. Peng, W.J. Peng, Y.W. Yu, G.X. Wu, J.Y. Zhang, Isothermal hydrogen
reduction of oxide scale on hot-rolled steel strip in 30 pct H2–N2 atmosphere,
Int. J. Hydrogen Energy 42 (50) (2017) 29921–29928.
[1] A.R. Marder, The metallurgy of zinc-coated steel, Prog. Mater Sci. 45 (3) (2000)
[25] A.R.P. Ghuman, J.I. Goldstein, Reaction mechanisms for the coatings formed
191–271.
during the hot dipping of iron in 0 to 10 Pct Al-Zn baths at 450° to 700°C,
[2] E. Baril, G. L’Espérance, Studies of the morphology of the Al-rich interfacial
Metall. Trans. 2 (10) (1971) 2903–2914.
layer formed during the hot dip galvanizing of steel sheet, Metall. Mater. Trans.
[26] F. Ajersch, F. Ilinca, J.-F. Hétu, Simulation of flow in a continuous galvanizing
A 30 (13) (1999) 681–695.
bath: Part I. Thermal effects of ingot addition, Metall. Mater. Trans. B 35 (1)
[3] J.R. McDermid, M.H. Kaye, W.T. Thompson, Fe solubility in the Zn-rich corner of
(2004) 161–170.
the Zn-Al-Fe system for use in continuous galvanizing and galvannealing,
[27] F. Ajersch, F. Ilinca, J.F. Hétu, Simulation of flow in a continuous galvanizing
Metall. Mater. Trans. B-Process Metall. Mater. Process. Sci. 38 (2) (2007) 215–
bath: Part II. Transient aluminum distribution resulting from ingot addition,
230.
Metall. Mater. Trans. B 35 (1) (2004) 171–178.
[4] L. Chen, R. Fourmentin, J.R. Mc Dermid, Morphology and kinetics of interfacial
[28] F. Ajersch, F. Ilinca, J.F. Hétu, F. Goodwin, Numerical simulation of flow,
layer formation during continuous hot-dip galvanizing and galvannealing,
temperature and composition variations in a galvanizing bath, Can. Metall. Q.
Metall. Mater. Trans. A-Phys. Metall. Mater. Sci. 39A (9) (2008) 2128–2142.
44 (3) (2005) 369–378.
[5] N.Y. Tang, Modeling Al enrichment in galvanized coatings, Metall. Mater.
[29] F. Ajersch, F. Ilinca, J.F. Hétu, F.E. Goodwin, Numerical simulation of the rate of
Trans. A 26 (7) (1995) 1699–1704.
dross formation in continuous galvanizing baths, Iron Steel Technol. 3 (8)
[6] P. Toussaint, L. Segers, R. Winand, M. Dubois, Mathematical modelling of Al
(2006) 93–101.
take-up during the interfacial inhibiting layer formation in continuous
[30] F. Ilinca, F. Ajersch, C. Baril, F.E. Goodwin, Numerical simulation of the
galvanizing, ISIJ Int. 38 (9) (1998) 985–990.
galvanizing process during GA to GI transition, Int. J. Numer. Meth. Fluids 53
[7] S. O’Dell, J. Charles, M. Vlot, V. Randle, Modelling of iron dissolution during hot
(10) (2007) 1629–1646.
dip galvanising of strip steel, Mater. Sci. Technol. 20 (2) (2004) 251–256.
[31] H.S. Park, K.A. Han, J. Lee, J.W. Shim, Numerical simulation of zinc flow and
[8] M.L. Giorgi, J.B. Guillot, R. Nicolle, Theoretical model of the interfacial reactions
temperature distribution in a galvanizing zinc pot, ISIJ Int. 48 (2) (2008) 224–
between solid iron and liquid zinc-aluminium alloy, J. Mater. Sci. 40 (9) (2005)
229.
2263–2268.
[32] T.P.C. Klaver, G.K.H. Madsen, R. Drautz, A DFT study of formation energies of
[9] M. Dutta, S.B. Singh, Effect of strip temperature on the formation of an Fe2Al5
Fe–Zn–Al intermetallics and solutes, Intermetallics 31 (2012) 137–144.
inhibition layer during hot-dip galvanizing, Scripta Mater. 60 (8) (2009) 643–
[33] C.H. Zhang, S. Huang, J. Shen, N.X. Chen, Structural and mechanical properties
646.
of Fe-Al compounds: an atomistic study by EAM simulation, Intermetallics 52
[10] G.K. Mandal, R. Balasubramaniam, S.P. Mehrotra, Theoretical investigation of
(2014) 86–91.
the interfacial reactions during hot-dip galvanizing of steel, Metall. Mater.
[34] T. Tsukahara, N. Takata, S. Kobayash, M. Takeyama, Mechanical properties of
Trans. A 40 (3) (2009) 637–645.
Fe2Al5 and FeAl3 intermetallic phases at ambient temperature, Tetsu To
[11] C.E. Jordan, A.R. Marder, Fe-Zn phase formation in interstitial-free steels hot-
Hagane-J, Iron Steel Inst. Jpn. 102 (2) (2016) 29–35.
dip galvanized at 450 °C: Part II 0.20 wt% Al-Zn baths, J. Mater. Sci. 32 (21)
[35] T. Zienert, L. Amirkhanyan, J. Seidel, R. Wirnata, T. Weissbach, T. Gruber, O.
(1997) 5603–5610.
Fabrichnaya, J. Kortus, Heat capacity of g-AlFe (Fe2Al5), Intermetallics 77
[12] C.E. Jordan, R. Zuhr, A.R. Marder, Effect of phosphorous surface segregation on
(2016) 14–22.
iron-zinc reaction kinetics during hot-dip galvanizing, Metall. Mater. Trans. A
[36] H. Becker, L. Amirkhanyan, J. Kortus, A. Leineweber, Powder-X-ray diffraction
28 (12) (1997) 2695–2703.
analysis of the crystal structure of the g0 -Al8Fe3 (g0 -Al2.67Fe) phase, J. Alloy.
[13] E. McDevitt, Y. Morimoto, M. Meshii, Characterization of the Fe-Al interfacial
Compd. 721 (2017) 691–696.
layer in a commercial hot-dip galvanized coating, ISIJ Int. 37 (8) (1997) 776–
[37] T. Zienert, A. Leineweber, O. Fabrichnaya, Heat capacity of Fe-Al intermetallics:
782.
B2-FeAl, FeAl2, Fe2Al5 and Fe4Al13, J. Alloy. Compd. 725 (2017) 848–859.
[14] Y. Morimoto, E. McDevitt, M. Meshii, Characterization of the Fe-Al inhibition
[38] S. Yang, X. Su, J. Wang, F. Yin, N.Y. Tang, Z. Li, X. Wang, Z. Zhu, H. Tu, X. Li,
layer formed in the initial stages of hot-dip galvannealing, ISIJ Int. 37 (9)
Comprehensive evaluation of aluminum diffusivity in liquid zinc, Metall.
(1997) 906–913.
Mater. Trans. A 42 (7) (2011) 1785–1792.
[15] K.K. Wang, L. Chang, D. Gan, H.P. Wang, Heteroepitaxial growth of Fe2Al5
[39] T. Min, Y.M. Gao, L. Chen, Q.J. Kang, W.Q. Tao, Mesoscale investigation of
inhibition layer in hot-dip galvanizing of an interstitial-free steel, Thin Solid
reaction-diffusion and structure evolution during Fe-Al inhibition layer
Films 518 (8) (2010) 1935–1942.
formation in hot-dip galvanizing, Int. J. Heat Mass Transf. 92 (2016) 370–380.
[16] R. Sagl, A. Jarosik, D. Stifter, G. Angeli, The role of surface oxides on annealed
[40] L. Chen, H.B. Luan, Y.L. He, W.Q. Tao, Pore-scale flow and mass transport in gas
high-strength steels in hot-dip galvanizing, Corros. Sci. 70 (Supplement C)
diffusion layer of proton exchange membrane fuel cell with interdigitated flow
(2013) 268–275.
fields, Int. J. Therm. Sci. 51 (2012) 132–144.
[17] K.K. Wang, C.W. Hsu, L. Chang, D. Gan, K.C. Yang, Role of Al in Zn bath on the
[41] L. Chen, Y. He, W.Q. Tao, P. Zelenay, R. Mukundan, Q. Kang, Pore-scale study of
formation of the inhibition layer during hot-dip galvanizing for a 1.2Si–1.5Mn
multiphase reactive transport in fibrous electrodes of vanadium redox flow
transformation-induced plasticity steel, Appl. Surf. Sci. 285 (Part B) (2013)
batteries, Electrochim. Acta 248 (2017) 425–439.
458–468.
[42] L. Chen, M. Wang, Q. Kang, W. Tao, Pore scale study of multiphase
[18] H. Yang, S. Zhang, J. Li, X. Liu, H. Wang, Effect of strip entry temperature on the
multicomponent reactive transport during CO2 dissolution trapping, Adv.
formation of interfacial layer during hot-dip galvanizing of press-hardened
Water Resour. 116 (2018) 208–218.
steel, Surf. Coat. Technol. 240 (Supplement C) (2014) 269–274.
[43] D.L. Beke, Y. Kaganovskii, G.L. Katona, Interdiffusion along grain boundaries –
[19] C. Guan, J. Li, N. Tan, S.G. Zhang, W.Y. Zhang, Effect of bath aluminum
diffusion Induced Grain Boundary Migration, low temperature
concentration on the galvanizing of hydrogen reduced hot rolled steel without
homogenization and reactions in nanostructured thin films, Prog. Mater Sci.
acid pickling, Surf. Coat. Technol. 279 (2015) 142–149.
(2018).
[20] N. Tan, J. Li, C. Guan, Investigation of hot rolled galvanised steel without acid
[44] T. Min, Y.M. Gao, L. Chen, Q.J. Kang, W.Q. Tao, Changes in porosity,
pickling, Ironmak. Steelmak. 40 (8) (2013) 578–581.
permeability and surface area during rock dissolution: effects of
[21] Z.F. Li, Y.Q. He, G.M. Gao, J.J. Tang, X.J. Zhang, Z.Y. Liu, Effects of Al contents on
mineralogical heterogeneity, Int. J. Heat Mass Transf. 103 (2016) 900–913.
microstructure and properties of hot-dip Zn-Al alloy coatings on hydrogen
[45] M. Blumenau, M. Norden, F. Friedel, K. Peters, Use of pre-oxidation to improve
reduced hot-rolled steel without acid pickling, J. Iron. Steel Res. Int. 24 (10)
reactive wetting of high manganese alloyed steel during hot-dip galvanizing,
(2017) 1032–1040.
Surf. Coat. Technol. 206 (2) (2011) 559–567.
[22] Y.Q. He, T. Jia, X.J. Liu, G.M. Cao, Z.Y. Liu, J. Li, Hot-dip galvanizing of carbon
[46] J.H. Park, G.H. Park, D.J. Paik, Y. Huh, M.H. Hong, Influence of aluminum on the
steel after cold rolling with oxide scale and hydrogen descaling, J. Iron Steel
formation behavior of Zn-Al-Fe intermetallic particles in a zinc bath, Metall.
Res. Int. 21 (2) (2014) 222–226.
Mater. Trans. A 43 (1) (2012) 195–207.
[23] C. Guan, J. Li, N. Tan, Y.Q. He, S.G. Zhang, Reduction of oxide scale on hot-rolled
[47] S. Feliu, V. Barranco, XPS study of the surface chemistry of conventional hot-
steel by hydrogen at low temperature, Int. J. Hydrogen Energy 39 (27) (2014)
dip galvanised pure Zn, galvanneal and Zn-Al alloy coatings on steel, Acta
15116–15124.
Mater. 51 (18) (2003) 5413–5424.

Das könnte Ihnen auch gefallen