ArticlePDF Available

Micro-Vickers Hardness of Intermetallic Compounds in the Zn-rich Portion of Zn–Fe Binary System

Authors:

Abstract and Figures

Annealing temperature and composition dependences of hardness for the intermetallic compound phases, especially Γ-Fe4Zn9, δ1p-Fe13Zn126 and δ1k-FeZn7, which are obtained in Zn-rich portion of Fe–Zn alloys, were investigated by the micro-Vickers hardness test. Although the hardness of the Γ and δ1k phases only slightly decreases with increasing quenching temperature, it shows obvious composition dependences. However, the hardness of the disordered δ1p phase is basically lower than that of the ordered δ1k phase, which is hardly affected by composition change. The hardness of all the Zn-rich compound phases quenched from 500°C (or 550°C) are evaluated and compared.
Content may be subject to copyright.
ISIJ International, Vol. 58 (2018), No. 9
© 2018 ISIJ 1578
ISIJ International, Vol. 58 (2018), No. 9, pp. 1578–1583
* Corresponding author: E-mail: k_han@material.tohoku.ac.jp
DOI: http://dx.doi.org/10.2355/isijinternational.ISIJINT-2018-111
1. Introduction
Hot-dip galvanizing (GI) and galvannealing (GA) tech-
niques are known to be very eective in improving the
corrosion resistance of steel sheets for automobile bodies
without change in material characteristics. The GI pro-
cess has been generally performed by the immersion of
steel plates into a molten Zn–Al bath maintained at about
445°C–455°C for several seconds,1) whereby very simple
coating layers,
η
Fe2Al5 (Cmcm) and Zn solid solution,
were mainly observed. Conversely, in the GA process, the
annealing is carried out at about 500°C1)–530°C2) after the
GI process; intermetallic compound (IMC) layers of the
Fe-rich portion in the Zn–Fe binary system are formed by
a long reaction time between the Fe substrate and
ζ
-FeZn13
(P63/mmc) layer.1,3–5) The processes are carefully controlled
to suppress defects such as aking and powdering that occur
in press forming.
Recently, the phase diagram of the Zn–Fe binary system
was experimentally determined by an alloying method,6)
and phase equilibria of ve IMCs, Γ-Fe4Zn9 (I
4
3m),
Γ1-Fe21.2Zn80.8 (F
4
3m),
δ
1k-FeZn7 (P63/mcm and R3c),
δ
1p-Fe13Zn126 (P63/mmc) and
ζ
were precisely conrmed as
shown in Fig. 1. The aking and the powdering are believed
to be directly related to formation of the Γ phase layer in the
diusion zone, because the fractures in the plating layer dur-
ing the deformation process are frequently observed on the
α
Fe/Γ phase boundary or in the Γ phase layer.1,7) Therefore,
mechanical properties of the IMCs (especially, Γ phase) are
a key issue to be examined for improving the formability of
Micro-Vickers Hardness of Intermetallic Compounds in the Zn-rich
Portion of Zn–Fe Binary System
Kwangsik HAN,1)* Inho LEE,1) Ikuo OHNUMA,2) Kaneharu OKUDA3) and Ryosuke KAINUMA1)
1) Department of Materials Science, Tohoku University, 6-6-02 Aobayama, Sendai, 980-8579 Japan.
2) Computational Materials Science Unit, National Institute for Materials Science (NIMS), 1-2-1 Sengen, Tsukuba, 305-0047
Japan. 3) Steel Research Laboratory, JFE Steel Corporation, Kawasaki, Japan.
(Received on February 19, 2018; accepted on March 20, 2018)
Annealing temperature and composition dependences of hardness for the intermetallic compound
phases, especially Γ-Fe4Zn9,
δ
1p-Fe13Zn126 and
δ
1k-FeZn7, which are obtained in Zn-rich portion of Fe–Zn
alloys, were investigated by the micro-Vickers hardness test. Although the hardness of the Γ and
δ
1k
phases only slightly decreases with increasing quenching temperature, it shows obvious composition
dependences. However, the hardness of the disordered
δ
1p phase is basically lower than that of the
ordered
δ
1k phase, which is hardly affected by composition change. The hardness of all the Zn-rich com-
pound phases quenched from 500°C (or 550°C) are evaluated and compared.
KEY WORDS: hot-dip galvanizing; vickers hardness; intermetallic compounds; annealing temperature and
composition dependences.
Zn-coated steel sheets.
The deformability of each IMC in the Fe–Zn system was
recently evaluated by Okamoto et al.7) by using a compres-
sion test of micropillars. According to them, the Γ phase has
enough ductility both in poly- and single-crystalline struc-
tures, satisfying von Mises criterion. Moreover, although the
ζ
phase is also ductile and shows a large plastic strain only
in single-crystal, the other three phases are basically brittle
in both poly- and single-crystal.
However, Vickers hardness test is also one of the most
important methods for most easily evaluating the mechani-
cal property of materials. Hardness of Γ, Γ1,
δ
and
ζ
have
already been reported by Bastin et al.,8) where the micro-
Vickers hardness test was performed for each IMC layer
Fig. 1. Phase diagram of the Zn–Fe binary system.6)
ISIJ International, Vol. 58 (2018), No. 9
© 2018 ISIJ1579
in a diusion couple (DC) specimen. According to them,
among the four IMCs, the highest and lowest hardness val-
ues were obtained in the Γ1 and
ζ
phases, respectively, and
the hardness of the
δ
phase is slightly higher than that of
the Γ phase. To precisely clarify the powdering mechanism,
the chemical-composition dependence of hardness may have
to be considered, especially in stoichiometric composition
(SC) and its compositional deviation. However, they have
reported no information on the composition dependence.
Recently, Kainuma et al.9) reported that while the second
order order-disorder transition from the disordered
δ
1p to the
ordered
δ
1k phase10,11) exists at temperatures above 550°C,
the ordering reaction changes to the rst order transition and
a phase separation (i.e.,
δ
1k +
δ
1p two-phase region) appears
at temperatures below 550°C. Very recently, the miscibil-
ity gap with a narrow two-phase region of only 0.5 at.%
has been conrmed by the present authors with two-phase
microstructure obtained by the alloying method.6) Further-
more, it has also been reported that the hardness of the
δ
1k
phase is much higher than that of the
δ
1p one and that the
dierence in hardness is available for identifying the struc-
ture.6) Eects of the chemical composition and the degree
of order on hardness of the ordered-
δ
1k phase are important
and interesting, but no systematic investigation on the issue
has yet been reported.
In the present study, dependence of hardness on anneal-
ing temperatures and compositions were examined by the
micro-Vickers hardness test. Especially, composition depen-
dences for Γ, ordered-
δ
1k and disordered-
δ
1p were precisely
investigated.
2. Experimental Procedures
The Fe–Zn alloys were prepared by high purity Fe (99.9
at.%) and Zn (99.99 at.%). Samples were melted by two-
step melting (TSM) and re-melting (RM) methods according
to the same procedure that was used in a previous study.6)
Prepared samples were equilibrated at temperature ranges
between 300°C and 700°C for various annealing times.
After equilibration, the hardness of each phase was very
carefully measured by a Micro-Vickers hardness tester
(Akashi, MVK-H1) under an applied load of 10 g (Fload =
0.01 × 9.8 N) for 10 seconds in two-phase microstructures.
To remove the inuence on the working hardening, the dis-
tance of each diamond indentation was located suciently
far from the previously measured location, as shown in Fig.
2. Sizes of indentations and horizontal (D1) and vertical (D2)
lengths were measured by observation of the eld-emission
scanning electron microscope (FE-SEM) at high magnica-
tion. The hardness was calculated by the following standard
formula:
HV F
DD
load



1 8544 0 102
2
12
2
..
/
,
where the units of Fload and Di are N and mm. The chemi-
cal compositions in the vicinity of the diamond indentation
were measured by wavelength dispersive X-ray spectros-
copy (WDS) equipped with a eld-emission electron probe
microanalyzer (FE-EPMA: JEOL JXA-8500F) under the
condition of an accelerating voltage of 20 kV and a beam
current of 10 nA. Transmission electron microscope (TEM)
samples were prepared by focused ion beam (FIB), and
selected area electron diraction (SAED) patterns were
observed by a scanning transmission electron microscope
(STEM: JEOL JEM ARM-200F) under the condition of an
accelerating voltage of 200 kV. The stable phases obtained
in the present study were plotted on the phase diagram as
shown in Fig. 3, and annealing temperatures, hardness and
chemical compositions are summarized in Table 1.
3. Results and Discussion
3.1. Microstructure Near Indentation and Hardness in
IMCs
Figure 4 shows the typical microstructure of diamond
indentation for every IMC phase in the
α
Fe , Γ+Γ
1,
Γ1+
δ
1k,
δ
1k +
δ
1p,
δ
1p +
ζ
and
ζ
+L(
η
) two-phase alloys
obtained by annealing at 500°C, where a pair of indenta-
tions, i.e., on the Zn-rich and Fe-rich sides, is presented
for each IMC. Note that those for the Zn-rich
δ
1k and the
Fe-rich
δ
1p were obtained from samples after heat treatment
at 550°C, because two-phases microstructures obtained at
Fig. 2. Typical BSE image of the diamond indentation for the Γ
and
δ
1k two phases after heat treatment at 600°C.
Fig. 3. The equilibrium compositions of samples for measurement
of hardness.
ISIJ International, Vol. 58 (2018), No. 9
© 2018 ISIJ 1580
500°C were too ne to be quantitatively evaluated even in
a hardness test with the lightest load. The peelings were
mostly observed in Γ1,
δ
1k,
δ
1p and
ζ
phases; the hairline
cracks mainly occurred in Fe-enriched
δ
1k, whereas no
defects were observed in Γ. In general, surface peelings or
hairline cracks are usually observed in brittle phases. Thus,
the microstructures in the diamond indentation suggest that
although most IMCs, especially in the
δ
1k phase on the Fe-
rich side, are very brittle, the Γ phase possesses a relatively
high ductility. A similar tendency has been reported by
Okamoto et al.7) using the micro-pillar compression test.
The hardness values of all the IMCs evaluated from the
indentations in Fig. 4 are shown in Fig. 5. The hardness val-
ues of the IMCs are observed to basically follow the order
of Γ1 >
δ
1k > Γ
δ
1p >
ζ
, which is the same tendency as
that of our previous report obtained from 450°C,12) whereas
the composition dependence is signicant in the Γ and the
δ
1k. The data reported by Bastin et al.8) from DC samples
heat treated at 450°C are also plotted in Fig. 5. While the
hardness data in the present study are slightly higher than
those on the whole, the tendency of the order in hardness
is almost consistent. The dierence in hardness from the
previous data may be due to the dierence of the loading
weight in the hardness test.13)
In the past, the GA process has been performed to
increase the Fe concentration in coating layers due to
decreasing friction coecient by formation of harder IMC
layers. Furthermore, the control of volume fraction of the
IMC layers with various Fe concentrations is one of the
most important issues to improve the press-formability.
The powdering in the GA process is known to mostly occur
when the Fe concentration in the coating layers is too high,
which may be consistent with the brittleness and the high
hardness obtained in the
δ
1k phase on the Fe-rich side. This
suggests that, to inhibit the powdering defect, the control
of Fe concentration in the
δ
1k is important. In the following
sections, the annealing temperature and chemical composi-
tion dependencies of hardness for the Γ,
δ
1k and
δ
1p phases
are presented.
3.2. Annealing Temperature Dependence of Hardness
for Γ,
δ
1k and
δ
1p
As shown in the phase diagram of Fig. 3, the phase exist-
ing in a wide temperature range is limited to the Γ and
δ
,
Fig. 4. Typical microstr uctures defor med by diamond indenta-
tion. (a, b) Γ-Fe4Zn9 (c, d) Γ1-Fe21.2Zn80.8 (e, f)
δ
1k-FeZn7 (g,
h)
δ
1p-Fe13Zn123 (i, j)
ζ
-FeZn13.
Tab le 1. Sum mary of annealing temperat ures, hardness and equi-
libria composition s.
Phase Te m p e r at u r e
(℃)
Hardness
(HV)
Equilibrium
composition
(at .%Zn)
Fe-enr iched Γ ()
300 393 ( ±13) 69.5 ( ±0. 3)
400 408 ( ±10) 69.1 ( ±0.2)
500 386 ( ±11) 68.5 (±0. 3)
600 387 ( ±17 ) 68.2 ( ±0. 3)
700 383 ( ±12) 68.4 ( ±0.2)
Zn-en riched Γ ()
500 440 ( ±10) 74.1 ( ± 0.1)
600 458 ( ±10) 75.7 ( ±0. 2)
650 475 ( ±14) 77.2 ( ±0.1)
675 481 ( ±13) 7 7.5 ( ±0.2)
700 460 ( ±6) 75.8 ( ±0. 2)
730 433 ( ±10) 73.5 ( ±0.2)
Fe-enr iched Γ1 ()500 587 ( ±9) 77.7 ( ±0.2)
Zn-en riched Γ1 ()500 560 ( ±15) 78.8 (± 0.1)
Fe-enr iched
δ
1k ()
500 525 ( ±12) 85.3 ( ±0.1)
600 505 ( ±6) 85.1 ( ±0.1)
650 502 ( ±14) 85.6 ( ±0. 2)
Zn-en riched
δ
1k ()
400 392 ( ±7) 90.8 (±0.0)
550 433 ( ±14) 87.8 ( ±0.1)
580 430 ( ±13) 89.3 ( ±0.1)
615 448 ( ±8) 86.5 ( ±0.1)
δ
1p ()
500 379 ( ±10) 91.5 ( ±0.1)
550 394 ( ±8) 90.1 ( ±0.1)
580 374 ( ±11) 90.3 ( ±0.1)
Fe-enr iched
ζ
()
400 298 ( ±7) 92.2 ( ±0.1)
500 280 ( ±16) 92.2 ( ±0. 2)
Zn-en riched
ζ
()500 291 ( ±10) 92.5 (±0.1)
ISIJ International, Vol. 58 (2018), No. 9
© 2018 ISIJ1581
and the dependence of annealing temperatures was evalu-
ated for the Γ,
δ
1k and
δ
1p, where the groups of hardness
data with similar composition were selected for each IMC.
Figure 6 shows the variation of hardness on the annealing
temperatures for the Γ,
δ
1k and
δ
1p phases with 68.8 ± 0.5,
85.4 ± 0.3 and 90.6 ± 0.6 at.% Zn, respectively. The hard-
ness is basically hardly aected by the annealing tempera-
ture, but all the IMCs actually decrease very faintly with
increasing annealing temperature. Generally, the Vickers
hardness depends on the amount of dislocations gener-
ated by mechanical deformations, and the atomic bonding
strength. Therefore, if neither phase transformation nor
change of degree of atomic order exists in the temperature
region, hardness at room temperature may be almost con-
stant, regardless of the annealing temperature. The reason
why the hardness for the IMCs slightly decreases at higher
temperatures is unclear.
3.3. Composition Dependence of Hardness for
Γ-Fe3Zn10,
δ
1k-FeZn7 and
δ
1p-Fe13Zn126
In Section 3.2, it was concluded that the hardness of
all the Γ,
δ
1k and
δ
1p phases are almost independent of
the annealing temperature. Therefore, we assume that the
tendency basically meets that of other alloy compositions
for each IMC, and estimate the composition dependence of
hardness by using the data obtained from the IMC phase
annealed even at dierent temperatures.
3.3.1. Hardness in the Γ Phase
Figure 7 shows hardness of composition dependence in
the Γ phase, together with the SC reported by Brandon et
al.,14) Johannson et al.15) and Okamoto et al.16) Compared
with the hardness in the alloys annealed at dierent tem-
peratures, it is conrmed that while the hardness of Γ hardly
depends on annealing temperature also in the Zn-rich por-
tion, it gradually increases with increasing Zn composition.
The relationship between composition and hardness is well
known and has been discussed in several papers.17–19) For
instance, minimum hardness might be indicated in SC due to
the smallest lattice defects such as vacancies, anti-site atoms
Fig. 5. Hardness of IMCs after heat treatment at 500°C (and 550°C) for Fe-en riched and Zn-enriched IMCs, together
with those at 450°C repor ted by Bastin et al.8)
and stacking faults and the hardness is gradually increased
with the compositional deviation from the SC. If this meets
the present case, the result shown in Fig. 7 means that the
SC of the Γ phase is not Fe3Zn10 of about 77.0 at.%Zn, at
least less than 69.5 at.%Zn.
The SC of Γ-Fe3Zn10 was reported to be 76.9 at.%Zn
(
γ
-brass structure) by Brandon et al.,12) and accepted by
Fi g. 7. Composition dependence of hardness for Γ-Fe4Zn9.
Fig. 6. Annealing temperat ure dependence of hardness for Fe-
enriched Γ-Fe4Zn9,
δ
1k-FeZn7 and
δ
1p-Fe13Zn126.
ISIJ International, Vol. 58 (2018), No. 9
© 2018 ISIJ 1582
Burton and Perrot.20) However, the stoichiometric ratio (SR)
of Γ-Fe3Zn10, NFe : NZn = 3 : 10, seems doubtful, because
the SC was determined by the results based on the XRD
examination, by which it is dicult to determine the exact
occupation sites of each atom in sub-lattices due to the
similar atomic scattering factors of Fe and Zn. Conversely,
several
γ
-brass type phases such as X3Y10, X4Y9 and X5Y8
are reported to form from the BCC or B2 closed structures
by a 26-atom cluster or two dierent 26-atom clusters.16,21,22)
The cluster has four types of sub-lattices, and the SC can be
varied by the element’s occupation. According to Gourdon
et al.,21) the sub-lattice of the Cu5Zn8 phase consists of an
outer-tetrahedron (OT: 4 atoms), an octahedron (OH: 6),
an inner-tetrahedron (IT: 4) and a distorted cuboctahedron
(CO: 12). The mainly occupied element is the Cu atom
in OT and OH, and the Zn atom in IT and CO. In these
sub-lattices, if the Cu site in the OH sub-lattice is partially
substituted by excess Zn atoms, the SC can be shifted from
Cu5Zn8 (61.5 at.%Zn) to Cu4Zn9 (69.2 at.%Zn). However,
very recently, Okamoto et al.16) re-examined the occupation
of Fe and Zn atoms in four sub-lattices and proposed an
atomic conguration model, which is based on the Fe+12Zn
icosahedron cluster structure. According to them, the Fe
atoms are occupied both in OT and IT sub-lattices, whereas
the OH and CO sub-lattices are fully occupied by the Zn
atom, as shown in Fig. 8, which is dierent from those of
Cu and Zn in the Cu5Zn8 phase. If all atoms fully occupy
each corresponding site, the SR of the Γ phase becomes
Fe4Zn9. The composition dependence of hardness in the
present study is consistent with that for SC with Fe4Zn9 of
69.2 at.%Zn, which was reported by Johannson et al.15) and
Okamoto et al.16)
3.3.2. Hardness in Ordered-
δ
1k and Disordered-
δ
1p Phases
The composition dependences of hardness on both
ordered-
δ
1k and disordered-
δ
1p phases are shown in Fig. 9,
together with the data in our previous paper.6) Owing to the
relationship between SC and hardness mentioned in Section
3.3.1, both hardness of the SC
δ
1k-FeZn7 at 87.5 at.%Zn
and
δ
1p-Fe13Zn126 at 90.6 at.%Zn were also expected to be
minimum and to indicate a “V shape” around the SC. In the
δ
1p, the hardness is observed to be almost constant at about
390 Hv against composition change, where the disordered
structure of the sample was conrmed from the SAED pat-
tern of [0001] incident direction taken from the 89.3 at.%Zn
sample annealed at 615°C, as shown in Fig. 10(b). If the
δ
1p phase obeys the general rule in hardness of IMCs, the
hardness has to increase with increasing Fe composition.
Fig. 8. Schematic illustration of the
γ
-brass cluster.16,21,22)
Fig. 9. Hardness distribution of ordered-
δ
1k and disordered-
δ
1p in
accordance with Zn concentration.
ISIJ International, Vol. 58 (2018), No. 9
© 2018 ISIJ1583
The constant hardness in the wide o-stoichiometric com-
position range may be brought about by the yielding stress
invariable for substitution of Fe atoms to some Zn sites.
The origin is not clear. Further studies on the substitutional
atomic sites of excess Fe atoms and the slip systems are
required. By contrast, the composition dependence in the
δ
1k is very complicated, i.e., the hardness almost linearly
decreases with increasing Zn composition to the SC com-
position of FeZn7 (87.5 at.%Zn), but from the SC composi-
tion suddenly changes to be almost constant and then drops
again to the level of hardness in the
δ
1p phase at about
90.4 at.%Zn. Here, the ordered conguration of the
δ
1k in
the 89.3 at.%Zn sample annealed at 580°C has been con-
rmed from the SAED pattern with the ordered spots of 1/3
{11
2
0}
δ
1p, as shown in Fig. 10(a). Thus, clearly, the hard-
ness of the ordered
δ
1k is higher than that of the disordered
δ
1p, as reported in our previous paper.6) Because the
δ
1p
δ
1k order-disorder transformation temperature drastically
decreases with increasing Zn composition, as shown in Fig.
9, the complicated behavior in the composition region over
the SC composition (FeZn7) in the
δ
1k may be explained by
the decrease of the degree of order. Consequently, the
δ
1k
at 90.4 at.%Zn should have a very low degree of order and
the hardness may be close to that of the
δ
1p.
All these results suggest that when controlling the micro-
structure in the coating layer including the
δ
phase, one
should take into account not only the kind of ordered or dis-
ordered phase, but also the Fe composition, in the
δ
phase.
4. Conclusions
The microstructures of indention and the hardness in the
IMCs appearing in Zn-rich portion of the Zn–Fe binary sys-
tem were examined by micro-Vickers hardness test.
(1) From the microstructures of indention, the peelings
were mostly observed in the Γ1,
δ
1k,
δ
1p and
ζ
phases and
the hairline cracks mainly occurred in the Fe-enriched
δ
1k,
whereas no defects were observed in Γ.
(2) The hardness values of the IMCs are basically in
order of Γ1 >
δ
1k > Γ
δ
1p >
ζ
.
(3) The hardness of the Γ,
δ
1k and
δ
1p phases were
almost independent of the annealing temperatures.
(4) In the Γ phase, the hardness gradually increases
with increasing Zn composition from 67.6 at.%Zn to 77.5
at.%Zn. This suggests that the SC of the Γ may not be at
Fe3Zn10.
(5) The hardness of
δ
1p is almost constant against the
Zn composition. However, the hardness of
δ
1k decreases
with increasing Zn composition and reaches an almost
similar level to that in disordered-
δ
1p on the Zn-rich side.
This means that the hardness of
δ
1k is strongly aected by
degree of order.
Acknowledgments
The authors would like to thank Prof. N. L. Okamoto at
Institute for Materials Research (IMR) of Tohoku university
for very meaningful discussion. This work was supported
by JSPS KAKENHI Grant Numbers 15H05766 and by The
Iron and Steel Institute of Japan.
REFERENCES
1) A. R. Marder: Prog. Mater. Sci., 45 (2000), 191.
2) J. Mackowial and N. R. Short: Int. Met. Rev., 24 (1979), 1.
3) C. E. Jordan and A. R. Marder: J. Mater. Sci., 32 (1997), 5593.
4) R. Kainuma and K. Ishida: ISIJ Int., 47 (2007), 740.
5) M. Safaeirad, M. R. Toroghinejad and F. Ashrazadeh: J. Mater.
Process. Technol., 196 (2008), 205.
6) K. Han, I. Ohnuma, K. Okuda and R. Kainuma: J. Alloy. Compd.,
737 (2018), 490.
7) N. L. Okamoto, D. Kashioka, M. Inomoto, H. Inui, H. Takebayashi
and S. Yamaguchi: Scr. Mater., 69 (2013), 307.
8) G. F. Bastin, F. J. J. van Loo and G. D. Rieck: Z. Metallkd., 65
(1974), 656.
9) R. Kainuma and K. Ishida: Tetsu-to-Hagané, 91 (2005), 349.
10) M. H. Hong and H. Saka: Scr. Mater., 36 (1997), 1423.
11) N. L. Okamoto, K. Tanaka, A. Yasuhara and H. Inui: Acta Crystal-
logr. B, 70 (2014), 275.
12) K. Han, I. Ohnuma, K. Okuda and R. Kainuma: Proc. 11th Int. Conf.
on Zinc and Zinc Alloy Coated Steel Sheet, (Galvatech 2017), ISIJ,
Tokyo, (2017), 83.
13) W. D. Nix and H. Gao: J. Mech. Phys. Solids, 46 (1998), 411.
14) J. K. Brandon, R. Y. Brizard, P. C. Chieh, R. K. Mcmillan and W.
B. Pearson: Acta Crystallogr. B, 30 (1974), 1412.
15) A. Johannsson, H. Ljung and S. Westman: Acta Chem. Scand., 22
(1968), 2743.
16) N. L. Okamoto, M. Inomoto, H. Takebayashi and H. Inui: J. Alloy.
Compd., 732 (2018), 52.
17) J. H. Westbrook: J. Electrochem. Soc., 103 (1956), 54.
18) L. M. Pike, Y. A. Chang and C. T. Liu: Acta Mater., 45 (1997), 3709.
19) J. H. Zhu, L. M. Pike, C. T. Liu and P. K. Liaw: Acta Mater., 47
(1999), 2003.
20) B. P. Burton and P. Perrot: Phase Diagram of Binary Iron Alloys, ed.
by H. Okamoto, ASM International, Material Park, OH, (1993), 459.
21) O. Gourdon, D. Gout, D. J. Williams, T. Proen, S. Hobbs and G. J.
Miller: Inorg. Chem., 46 (2007), 251.
22) T. B. Massalski: Mater. Trans., 51 (2010), 583.
Fig. 10. Typical SAED patterns obtained from [0001] direction of
(a) ordered-
δ
1k and (b) disordered-
δ
1p. Each of the phases
were obtained from the annealed samples of A and B in
Fig . 9.
... According to pertinent research, it has been established that for hot-formed sheet metal during the forming process, equations representing force balance, energy balance, and volume balance can be formulated [25][26][27] (Equations (1)-(3)). When these equations are integrated with the specific parameters associated with galvanized boron steel sheet materials [28][29][30][31][32][33], it becomes feasible to compute the real contact area, denoted as Ar. This calculation is achieved by normalizing the micro-asperities present on the surface of the sheet metal, thereby enabling the determination of the real contact area under various wear conditions. ...
Article
Full-text available
In the hot stamping process, the friction and wear interaction between the high-temperature sheet metal and the water-cooled die has a significant impact on the final quality of the product and the durability of the die. Currently, most research on the wear of the stamped parts during the hot stamping process mainly involves analyzing the wear morphology and wear mechanism of the sheet surface, and there is little research on its wear assessment. In this study, to better assess the forming quality of hot stamping parts, the research takes the direct hot stamping of galvanized ultra-high-strength steel sheets as the object and proposes a wear amount calculation method of galvanized ultra-high-strength steel sheets based on the real contact area of the high-temperature sheet metal and the water-cooled tools. At different temperature conditions, the galvanized layer and steel substrate have different mechanical properties. The model is validated using the sheet characteristics at 650 °C, 700 °C, and 750 °C. The results indicate that the model can predict the wear of the galvanized steel sheet under different conditions within a certain range.
... Meanwhile, when the Fe content in the coating layer increases (Fe-Zn alloying proceeds by varying the annealing time and/or temperature), the fraction of the relatively brittle phase increases [2]. The increased fraction of Fe in the phase produces a more brittle phase, which aggravates the powdering resistance [14]. In addition, the Γ phase also becomes consistently thicker under increasing Fe content in the coating layer. ...
Article
Full-text available
The failure of galvannealed (GA) coatings during press forming is an important issue for steel companies, because it results in a deteriorated product quality and reduced productivity. Powdering and flaking are thought to be the main failure modes in GA steel. However, these two modes currently lack a clear distinction, despite their different failure types. Therefore, in this study, we demonstrate that the different behaviors of these two failure modes are generated by the skin pass mill (SPM) condition and we discuss the underlying mechanism in detail using microstructural and simulation analyses. With the increase in steel elongation from 0% to 4.0% under milling force from 0 to 6 ton, a high compressive stress is produced up to −380 MPa on the surface of the steel sheet and the interface is correspondingly flattened from 0.96 to 0.53 m in Ra. This flattening weakens the mechanical interlocking effect for adhesive bonding, deteriorating the flaking resistance from 41.1 to 65.2 hat-bead contrast index (hci). In addition, the GA coating layer becomes uniformly densified via the filling of pores under compressive stress in the layer. Furthermore, the ζ phase exhibits significant plastic deformation, leading to a uniform coverage of the coating surface; this helps to suppress crack propagation. Accordingly, the powdering resistance gradually improves from 4.2 to 3.5 mm. Consequently, with the increase in SPM-realized steel sheet elongation, the powdering resistance improves whilst the flaking resistance deteriorates. Significantly for the literature, this implies that the two failure modes occur via different mechanisms and it indicates the possibility of controlling the two coating failure modes via the SPM conditions.
... The hardness was measured for 21 points under an applied load of 50 g for 12 s by the micro-Vickers hardness tester (Mitutoyo, HM-100); 7 points were measured for three different locations: the top, middle and bottom of the sample. The details for the measurement of the horizontal and vertical lengths of the diamond indentations and the hardness calculation method were described in a previous study [50]. The chemical compositions of each point were confirmed by the results of the WDS analysis obtained at approximately 5 µm from each diamond indentation, and the sample composition was averaged. ...
... Conversely, the δ 1k -FeZn 7 and δ 1p -Fe 13 Zn 126 phases, which had been expected to have ductility, were verified to be brittle, similar to the Γ 1 -Fe 21.2 Zn 80.8 phase. Furthermore, the δ 1k phase, composed of the Fe-rich side, was confirmed to be the most brittle phase, as evidenced by the edge cracks at the diamond indentations [6]. These results indicate that the formation and growth of desirable IMC layers, as well as the concentrations of Fe and Zn in each layer, should be optimized to prevent interfacial fracture. ...
Article
In the present study, the formation and growth behaviors of intermetallic compound (IMC) layers in the interfacial reaction of solid Fe and molten Zn during hot dipping at 450 °C were investigated. In the early stage of the interfacial reaction, Γ-Fe4Zn9, δ1k-FeZn7, and ζ-FeZn13 phases were formed, among which the growth of the ζ layer was dominant. Columns of the δ1p-Fe13Zn126 phase were formed on the δ1k/ζ interface at approximately 90 s, and the four IMC layers grew simultaneously until 600 s. Subsequently, a layer of the Γ1-Fe21.2Zn80.8 phase was formed along the Γ/δ1k interface and started growing, after which all the equilibrium IMC phases of Γ/Γ1/δ1k/δ1p/ζ grew competitively. The total thickness of the five IMC layers, dtotal, increased proportionally with the square root of the reaction time, t, i.e., dtotal = d0 t 0.5, which suggested that the growth of the entire IMC layers was controlled by the volume diffusion. However, the values of the time exponent, n, for the individual IMC layers differed depending on their microstructures and chemical concentrations, under equilibrium or nonequilibrium conditions. The formation of the multilayered columnar ζ phase can be attributed to the heterogeneous supersaturation of Fe in the liquid Zn near the columnar ζ / liquid Zn interface during the early stage of the interfacial reaction. The interdiffusion fluxes, J̃ϕ, of Fe and Zn in the ϕ phase (ϕ =Γ, δ1k, and ζ) were estimated from the measured concentration profiles of each phase. This suggested that J̃ϕ caused the growth of each phase as well as the supersaturation of Fe near the interfacial boundaries, thereby resulting in considerable deviations from the equilibrium concentrations. In the later stage of the interfacial reaction, the interfacial concentrations gradually approached the equilibrium concentrations with decreasing J̃ϕ.
... The Vickers hardness of the ζ phase in our study accorded with what Han et al.22) did, however, they differed with those of Han et al. with regards to the δ 1 phase. In our study, the Zn content of the δ 1 phase was measured to be 87.9%. ...
Article
Full-text available
To understand the fatigue mechanism of hot-dip galvanized steel, the fatigue strength and fracture surface of hot-dip galvanized AISI 1045 steel(carbon steel) specimens were investigated. The galvanized coating layer was composed of δ1-phase, ζ-phase and η-phase, and its thickness was about 100 µm. In the low cycle region (104 cycles < Nf < 105 cycles), the fatigue strengths of both the carbon steel and the galvanized steel corresponded with the static strength. The fatigue strength of the galvanized steel was lower than that of carbon steel. As the number of cycles increased, the difference between fatigue strength of the carbon steel and that of the galvanized steel increased. Also, the morphologies of the fatigue fracture were different in low cycle region and high cycle region. In the galvanized steel, the morphology of Stage II crack on the fracture surface at low cycle region exhibited crescent shape, and multiple crack initiation sites in low cycle region were observed. Whereas the morphology at high cycle region (Nf >105 cycles) exhibited an ellipse shape, and the crack initiation site was single. At both regions, the crack initiation sites were in the coating layer. The mechanical properties of the microstructure in the coating layer had an effect on the fatigue strength. When η-phase was removed from the galvanized coating layer, the fatigue strength increased only in the high cycle region. Therefore, δ1-phase and/or ζ-phase cause the fatigue strength to decrease in low cycle region, and η-phase causes it in high cycle region.
... Γ-Fe 3 Zn 10 is an inherently hard and brittle intermetallic compound. Han et al. [27] reports that the hardness of Γ phase ranges from 383 to 481 HV and the deformability is 8 times smaller than iron [28], which deteriorates the ductility and induces crack initiation. Fig. 17e is a representation of the third stage when the local temperature drops below 665 • C, and a new type of Fe-Zn intermetallic compound (δ-FeZn 10 ) is formed from another peritectic reaction between Γ-Fe 3 Zn 10 and zinc. ...
Article
Liquid metal embrittlement (LME) was observed in the laser lap welds of zinc-coated GEN3 steels. When line energy input ranged from 54 kJ/m to 60 kJ/m, the LME crack initiated from the weld edge at the faying interface, propagated along the growth direction of columnar grains and extended upwards along the weld centerline, exhibiting an inversed Y-shape. Multiple Fe–Zn phases such as α-Fe (Zn), Γ-Fe3Zn10 and δ-FeZn10 were created by liquid zinc in grain boundaries reacting with the austenite matrix during the cooling stage of the laser welding process. Those hard and brittle Fe–Zn phases, especially Γ-Fe3Zn10, deteriorated the grain boundary cohesion and led to crack initiation. Simulation results indicate that a weld profile being deep with partial penetration will result in a high stress concentration at the notch root during the LME active temperature range. Therefore, full penetration welding is recommended to reduce the risk of LME by enabling zinc vapor to escape from both the top and bottom of the keyhole.
Article
Galvanizing is an important industrial process to improve the corrosion resistance of advanced high strength steels (AHSSs) that are vital for automotive industries. During galvanizing, nanoscale intermetallic phases with complex crystal structures are formed at the interface between the steel substrate and the zinc overlay. To better understand the nanoscale structures and the interfacial properties between the intermetallics, in this work, we develop a second nearest neighbor (2NN) Fe-Al-Zn ternary Modified Embedded Atom Method (MEAM) potential to describe the crystal structures of the intermetallics, i.e., Fe3Al8, Fe4Zn9 and FeZn13 and to calculate the interfacial structure and energy between them. The developed MEAM potential describes well the complex crystal structures and can be used to investigate the interfacial properties that are difficult to obtain from experiments. The Fe4Zn9, FeZn13 surface energies; the Fe-Fe4Zn9, Fe-FeZn13, Fe3Al8-FeZn13 interfacial energies; and the work of adhesion (WOA) are calculated with the developed MEAM potential. The results show that FeZn13 crystal orientation has an insignificant effect on the FeZn13 surface energy and the Fe-FeZn13 interfacial energy. A negative interfacial energy is obtained for the Fe-Fe4Zn9 and the Fe-FeZn13 interface. The lowest interfacial energy is obtained in the {100}Fe case. The interfacial energy of Fe3Al8-FeZn13 depends on the surface termination of Fe3Al8 and FeZn13. A low interfacial energy is obtained when the surface termination of Fe3Al8 and FeZn13 are both Fe rich. In contrast, when the surface termination of Fe3Al8 is Al rich or the surface termination of FeZn13 is Zn rich, no low energy, stable interface can be formed between the two phases.
Article
The phase diagram of the Fe-Zn binary system was evaluated based on the CALPHAD method with reference to the latest experimental data. The solubility ranges of the intermetallic compound phases, Γ-Fe4Zn9, Γ1-Fe11Zn40, δ1k-FeZn7, δ1p- Fe13Zn126, and ζ-FeZn13 were modeled considering their structures consisting of Zn12 icosahedra with Fe at the center (Fe1Zn12 clusters) as well as glue-like Fe and Zn atoms, and the miscibility gap between the δ1k and δ1p phases was also taken into account in the present calculations. The solubility of Fe in the liquid and (ηZn) phases that was confirmed as dozens of times larger than the values reported in the earlier literature could be calculated by introducing Fe1Zn12 associates to these solution phases. Consequently, all phase equilibria were adequately reproduced by the thermodynamic models and parameters revised in the present study.
Article
Hot stamping of steel is an innovative process for producing components of higher strength with a significant weight reduction. Conventional hot stamping process is carried out at an elevated temperature in the range of 800 to 950 °C in atmospheric condition. So, a coating is a must to restrict high temperature oxidation and decarburization of the steel substrate. Majorly, Al-Si coating and Zn-based coatings are used for the purpose. However, Al-Si-based coatings are not able to provide sacrificial cathodic protection to steel substrate and the Zn-based coating suffer from the problem of microcracking. In the present study, the pre-Ni coated galvanized boron steels are investigated to understand high temperature phase transformation during austenitization step. Thermodynamic calculations along with Gleeble heat treatment experiments and coating characterizations with SEM, EDS, WDS and XRD shows initial melting of the coating. Subsequent Fe-enrichment in the coating has led to formation of BCC phase which is high melting as well as softer phase compared to Ni-Zn intermetallic. Moreover, the corrosion performance of the prior Ni coated galvanized steel has been evaluated. The presence of zinc in the coating shows sacrificial corrosion protection behavior and the Ni and Fe in the coating have further improved the corrosion rate owing to their barrier protection to the steel.
Article
Full-text available
Various experimental analyses on hydrogen evolution, absorption, and cracking behaviors were conducted to gain a fundamental understanding of the hydrogen embrittlement of ultrastrong steel sheets with galvanized (GI) and galvannealed (GA) coatings. The hydrogen evolution and absorption behaviors are controlled primarily by the potential differences between the coating and exposed steel substrate, and the corrosion-induced damage pattern of the coating. The higher absorption rate of hydrogen was more pronounced in corroded GI-coated steel caused by the larger cathodic polarization applied to the exposed substrate, and a more severe form of coating dissolution by aqueous corrosion in a 3.5% NaCl + 0.3% NH4SN solution. In contrast, the corrosive species can only penetrate through the pre-existing cracks in the brittle Fe-Zn intermetallic phases composed of the GA coating, and the driving force for hydrogen evolution becomes smaller. These result in significant differences in hydrogen penetration and cracking behaviors between the two coated ultrastrong steels.
Article
Full-text available
The crystal structures of the Γ- and Γ1-phase compounds in the Fe-Zn system have been refined by single-crystal synchrotron X-ray diffraction. The Γ-phase compound consists of four Fe-centred (Zn,Fe)12 icosahedra connected with one another by face-sharing with the vertex sites of the shared triangle face being occupied by both Fe and Zn atoms. The Fe occupancy at the vertex sites increases with increasing Fe composition of the compound. The stoichiometry of the Γ phase is described as Fe4Zn9 (Zn-30.77 at.%Fe), realized when all the vertex sites of the shared triangle face are occupied exclusively by Fe atoms. The Γ1-phase compound consists of Zn12 and (Zn,Fe)12 icosahedra whose central site is exclusively occupied by Fe atoms, which is different from the previously reported structural model where some centres of these icosahedra are occupied by both Fe and Zn atoms. The crystal structures of all Fe-Zn intermetallic compounds can thus be best understood by considering the packing of the common structural unit; Fe-centred Zn12 and/or Fe-centred (Zn,Fe)12 icosahedron. Although the crystal structures of the α-Fe, Γ and Γ1 phases with cubic symmetries are related with each other in terms of the arrangement of structural vacancies, there is no orientation relationship either between the Γ and Γ1 phases or between the α-Fe and Γ phases in the coating layer of galvannealed steel sheet.
Article
Full-text available
Micropillar compression tests on each of the five intermetallic phases of the Fe-Zn system, which constitute the coating of galvannealed steels, have revealed that the Gamma phase formed in direct contact with the steel substrate and the zeta phase formed on the outermost surface are ductile, sandwiching the other three brittle phases (Gamma(1), delta(1k) and delta(1p)). Compression deformability of these ductile phases is considered to mitigate the coating failure through sustaining ruptured fragments of the brittle phases during the forming process.
Article
Full-text available
The structure of the δ1p phase in the iron-zinc system has been refined by single-crystal synchrotron X-ray diffraction combined with scanning transmission electron microscopy. The large hexagonal unit cell of the δ1p phase with the space group of P63/mmc comprises more or less regular (normal) Zn12 icosahedra, disordered Zn12 icosahedra, Zn16 icosioctahedra and dangling Zn atoms that do not constitute any polyhedra. The unit cell contains 52 Fe and 504 Zn atoms so that the compound is expressed with the chemical formula of Fe13Zn126. All Fe atoms exclusively occupy the centre of normal and disordered icosahedra. Iron-centred normal icosahedra are linked to one another by face- and vertex-sharing forming two types of basal slabs, which are bridged with each other by face-sharing with icosioctahedra, whereas disordered icosahedra with positional disorder at their vertex sites are isolated from other polyhedra. The bonding features in the δ1p phase are discussed in comparison with those in the Γ and ζ phases in the iron-zinc system.
Article
A critical determination of a phase diagram of the Zn-Fe binary system was carried out by EPMA analysis for skillfully prepared and equilibrated two-phase alloys for solid phases and by EDS areal analysis for the Fe solubility in the heterogeneously solidified liquid phase instead of the conventional diffusion couple method. It was newly confirmed that equilibrium compositions of intermetallic compounds tend to shift toward the Fe side compared to the previous phase diagram on the whole and that the solubility ranges of the Γ-Fe3Zn10 and Γ1-Fe11Zn40 phases are much narrower than those in the literature. The critical composition of the second-order order-disorder transition between δ1k-FeZn7 and δ1p-FeZn10 at high temperature was estimated by Vickers hardness measurement. Furthermore, the phase separation, i.e., the first-order transition, between the δ1k and δ1p phases in the low temperature region was directly found out and its equilibrium compositions were confirmed by EPMA analysis for the two-phase microstructures. The width of the δ1k + δ1p two-phase region at 500 °C is only 0.5 at.% and the δ1p decomposes into the δ1k and ζ-FeZn13 phases via a eutectoid reaction at a temperature between 455 °C and 445 °C. While being mostly coincident with those in the previous literature at temperatures between 1000 °C and 600 °C, the solubility of Fe in the liquid (L) Zn phase is about twenty times larger than the assessed value in the literature at temperatures below 500 °C. A eutectic reaction, L → ζ + (ηZn), was confirmed to occur by precise EPMA analysis and DSC measurement.
Article
The microstructural evolution of intermetallic compound layers formed in Fe/Zn diffusion couples (DCs) was examined by optical and scanning electron microscopies and electron-probe microanalysis. In the solid-Fe/liquid-Zn DCs at 450°C, the δ p phase nucleates between the δ p and ζ phases after dipping for about 100 s and every intermediate phase seems to obey the square root law individually before and after the appearance of the δ p phase at 100 s. In the concentration-penetration profiles, a composition gap is observed in the δ p phase of the solid-Fe/liquid-Zn DCs dipped for a long time at temperatures ranging from 450 to 550°C, while only a singular point, but no gap, is obtained in the DCs dipped at elevated temperatures over 575°C. On the other hand, in the case of the solid-Fe/solid-Zn DCs at 400°C, the δ k phase instead of the δ p phase appeares later between the δ p and Γ phases and finally covers the whole δ region.
Article
By a diffusion couple technique, a new compound, called GAMMA //1, containing 18. 5 to 23. 5 at. % Fe at 380 degree C, was found in the iron zinc system. Its lattice is related to that of the GAMMA phase in that its cell parameter is obtained by doubling that of the bcc unit cell of the GAMMA phase, yielding a value of 17. 963 A. The new cell is of the fcc type. The homogeneity range of the GAMMA //1 phase was found to decrease with increasing temperature whereas the reverse was observed for the GAMMA phase.
Article
Reactive diffusion between solid Fe and liquid Zn was experimentally examined using Fe/Zn diffusion couples prepared by an immersion technique. Using this technique, a pure iron sheet was immersed in a molten pure Zn bath with a constant temperature of T=723 K (450 degrees C) for various times up to t=7.2x10(3)s (2 h). The microstructure of the cross section of the Fe/Zn diffusion couple was observed by scanning electron microscopy, and the chemical composition of each phase was determined by electron probe microanalysis. Interface concentrations of Fe for the zeta-FeZn13 and delta-FeZn7-10 phases at the zeta/delta interface, c(zeta delta) and c(delta zeta), and for the liquid-Zn (L) phase at the zeta/L interface, c(L zeta), determined from the diffusion couples were found to be extremely higher than the corresponding equilibrium values, especially in the early stage of t<300 s, and the difference gradually decreased during immersion. The deviations from the equilibrium values for the c(zeta delta) and CL were proportional to the interduffusion flux across the; phase layer.
Article
Several interesting features in the study of stabilities of phases, and in phase transformations, are discussed. It is proposed that symmetry considerations related to the presence of magnetism in iron suggests that the respective phases, BCC alpha and FCC gamma, have in fact lower symmetries than cubic. A proposal is made that the symbol beta used in the past for the designation of the paramagnetic BCC iron should perhaps be returned as a feature in phase diagrams. The importance of the new concept of a 'pseudogap' in the electronic band structure, as a stabilizing electronic feature, is discussed in the light of the Hume-Rothery electron concentration rule. It is proposed that since the thermal activation is a major feature in the behavior of isothermal martensites, a more suitable designation for these types of phase transformations might be "thermally activated martensites", or TAMs. Massive transformations are discussed briefly and it is emphasized that they present a specific example of an idiomorphic transformation process, not requiring the need for orientation relationships (ORs) between the parent and product phases. [doi:10.2320/matertrans.M2010012]