You are on page 1of 11

Surface & Coatings Technology 245 (2014) 55–65

Contents lists available at ScienceDirect

Surface & Coatings Technology


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

Effects of manganese addition on microstructures and corrosion behavior


of hot-dip zinc coatings of hot-rolled steels
Youbin Wang, Jianmin Zeng ⁎
State Key Laboratory of Solidification Processing, Northwestern Polytechnical University, Xi'an 710072, PR China

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

Article history: In the present work, the Zn coatings of hot-rolled steel sheets with different Mn additions were prepared by hot
Received 23 October 2013 dipping process. The effects of Mn addition on the microstructures and corrosion resistances of the hot-dip Zn
Accepted in revised form 19 February 2014 coatings were investigated by scanning electron microscopy (SEM), X-ray photoelectron spectroscopy (XPS),
Available online 26 February 2014
and X-ray diffractometry (XRD). The Mn distribution in δ-layer is lower than that of ζ-layer. Mn addition could
promote the growth of δ-layer of the coating and inhibit the growth of the columnar ζ phase which morphology
Keywords:
Hot dip
was changed to discontinuous phase. It is also found that the surface color of Zn–Mn coating varies with the
Zn coating dipping temperatures. Such phenomenon is supposed to be attributed to the optical interference effect of the
Mn addition Mn oxide film formed on the coating surface. For the corrosion behavior, the as-dipped and ζ layer-exposed Zn
Microstructure coatings were subjected to both Tafel polarization measurements and electrochemical impedance spectroscopy
Corrosion (EIS) tests in 3.5% NaCl solution. The results indicate that Mn addition enhances the corrosion resistance of
the as-dipped coatings. The dense corrosion products formed on the top η layer of the Zn–Mn coatings after Tafel
polarization corrosion are assumed to inhibit the further corrosion of the coating. However, minor effects of Mn
addition to the anti-corrosion performance of ζ layer were noticed.
© 2014 Elsevier B.V. All rights reserved.

1. Introduction [15,17–19]. However, due to prolonged processing time, high operating


cost, and limited work piece size and shape, the electrochemical deposi-
Hot-dip Zn coatings have been widely used for protection of steel tion process is not accepted widely in the industry. On the other hand,
work pieces against atmospheric corrosion owing to their excellent the hot dipping process would still be the primary choice for Zn–Mn
corrosion performance and low cost [1]. The coating typically consists coating preparation because of its low cost and easy operation in the
of multiple sub-layers including η-Zn layer, ζ-FeZn13 layer, δ-FeZn10 industry [1]. Many works have been carried out to investigate the
layer and Γ-Fe3Zn10 layer [1]. The enhancement of the corrosion resis- anti-corrosion performance of hot-dip Zn–Mn coating. For example,
tance of steel work pieces with Zn coatings is realized in two ways in E. Pavlidou [20] reported that 1.0% Mn addition improved the anti-
general, i.e., barrier protection to simply insulate the steel work pieces corrosion performance of the hot-dip Zn coating, where Mn acts as a
from corrosive environment, as well as galvanic protection to prevent sacrificial anode to protect Zn. L.D. Ballote [21] discovered that the addi-
the substrate steel from corrosion serving as sacrificial anode [1,2]. tion of Mn introduced Mn-rich oxides on the coating surface which
The overall performance of hot-dip Zn coating depends on many aspects therefore increased the corrosion resistance of hot-dip Zn coatings.
such as chemical compositions of steel work pieces [3], alloying element However, most of the previous researches were focused on the effect
contents in Zn bath [4], as well as dipping time and temperature [5]. of Mn addition on the corrosion resistance of the coating surface only.
Numerous works have been carried out in order to improve the corro- To the authors' best knowledge, the effect of Mn addition on the micro-
sion resistance of the coating by alloying Zn with small amounts of structure and corrosion behavior of the Zn–Fe sub-layers of the hot-dip
other elements, such as Al [6,7], Si [8,9], Ni [10], Ti [11], Mg [12], Co Zn coating are still not yet well studied. In practice, the Zn–Fe sub-layers
[13], and Sb [14], through better controlling the growth of the Zn–Fe would probably be exposed to the corrosion environments especially
phase, coating thickness, adhesion strength, Sandelin effect, and so on. when the coating has been subjected to some physical and/or chemical
Among the investigated Zn coatings, the Zn–Mn coating were of damages, such as accident peeling off of the top η-Zn layer during mov-
great interest due to the better anti-corrosion performance than ing, acid attack, etc. The Zn–Fe sub-layers would then be serving the
pure Zn coating [15,16]. Electrochemical deposition is one of effective important role in the protection of the substrates from corrosion. It is
means to produce the Zn–Mn coating with better corrosion resistance therefore worthwhile to investigate the corrosion resistance of such
sub-layers to further understand the anti-corrosion mechanisms of the
⁎ Corresponding author. Tel.: +86 771 3231238. Zn–Mn coatings. For this, the present work is trying to give more in-
E-mail address: zjmg2008@gmail.com (J. Zeng). sightful investigation on the microstructure and corrosion resistance

http://dx.doi.org/10.1016/j.surfcoat.2014.02.040
0257-8972/© 2014 Elsevier B.V. All rights reserved.
56 Y. Wang, J. Zeng / Surface & Coatings Technology 245 (2014) 55–65

of both the surface and the ζ-FeZn13 layer of the hot-dip Zn coating with investigation of the ζ layer's corrosion resistance. As we can see from
Mn addition. The effects of Mn addition on the corrosion behavior of Fig. 3, a 30–35 μm removal would be sufficient enough to expose the ζ
other Zn–Fe layers (Delta & Gamma) of the coating will be studied in fu- layer. In order to reach the desired ζ layer, a micrometer (FMS 283,
ture work. 1 μm resolution) was initially used to estimate the thickness of the coat-
ing being removed. After roughly 30 μm of coating being removed, SEM
2. Experimental and XRD were used to confirm the arrival of ζ layer, and the polishing
process was repeated until the ζ layer is completely exposed. All the
The Zn–Mn alloys with different Mn contents (0.2, 0.5 and 0.7 wt.%) samples were sealed by epoxy resin but leaving the testing surface
used for steel hot dip coating were prepared by adding certain amounts (1 cm2) uncovered for the corrosion tests. The electrochemical experi-
of as-purchased Zn–Mn intermediate alloy (5.0 wt.% Mn) into the mol- ments were performed in 3.5% NaCl solution at 20 °C with a classic
ten pure Zn (99.99 wt.%) which was previously melted at 650 °C in a three-electrode cell configuration in which a Pt plate served as
resistance furnace with a graphite crucible (25 cm in diameter and the counter electrode and a saturated calomel electrode (SCE) was
30 cm in depth). For the sake of simplicity, Zn–xMn is denoted for the used as the reference electrode. The polarization tests were carried
prepared Zn coating with x wt.% Mn addition throughout the whole ar- out at the scan rate of 0.2 mV/s and the scan range from − 200 mV
ticle. After sufficient mixing, the molten Zn–Mn alloy was hold at certain (SCE) to +200 mV (SCE) relative to the open-circuit potential. The elec-
temperatures, such as 430, 450, and 470 °C for steel dipping in the next trochemical impedance spectroscopy (EIS) measurements were con-
step. The chemical compositions of the prepared Zn alloys with different ducted at the potential amplitude of 10 mV relative to the open-circuit
Mn additions were determined by XRF (Olympus Innov-X) and listed in potential and the frequency from 10 mHz to 100 kHz.
Table 1.
Hot-rolled steel sheet (Grade Q235, containing 0.11 wt.% C, 0.55 wt.% 3. Results and discussion
Mn, 0.012 wt.% Si, 0.016 wt.% P and balance Fe, 3 mm in thickness) was
cut into specimens with the size of 90 mm (L) × 30 mm (W). Alkali–acid 3.1. Microstructure
wash was used to clean the surface of steel work pieces. The specimens
were first degreased in 10 wt.% NaOH solution for 20 min, followed by For comparison, the steel samples with hot-dip pure Zn and Zn–
rinsing in distilled water several times to completely remove the alkali 0.5Mn coatings prepared under different hot-dipping temperatures
solution. The specimens were then pickled in an acid solution with are presented in Fig. 1. Fig. 1(a) shows the hot-dip pure Zn coating
16 vol.% HCl for 20 min, followed by rinsing in distilled water multiple prepared at 450 °C in which the coating color appears gray white. No
times to remove the residual acid. The cleaned specimens were fluxed color variation was observed for the pure Zn coatings prepared at any
in an aqueous solution containing 15 wt.% ZnCl2 and 15 wt.% NH4Cl other hot-dipping temperatures. However, the colors were different
for 10 min, and then dried in a furnace at 120 °C. The flux coated on for those Zn–0.5Mn coatings prepared at different hot dipping temper-
the specimens helps prevent the steel surface from oxidation during atures. The surface color changes in a sequence of pale yellow, purple
the dipping process. Finally the specimens were dipped in the molten and cyan, as shown in Fig. 1(b–d), with the increase of the dipping
Zn bath hold in the resistance furnace. The dipping temperatures of temperature (430, 450 and 470 °C, respectively). It, to some extent, in-
the specimens were controlled at 430, 450, and 470 °C, respectively. dicates that the surface color of Zn–0.5Mn coating could be controlled
After 180 s of dipping, the specimens were removed from the molten by the dipping temperature. For other specimens with 0.2 wt.% and
Zn bath and cooled in air naturally. 0.7 wt.% Mn additions, same color variations were also observed.
The coating morphology was observed using a Hitachi S-3400N The color variation of the Zn–0.5Mn coating with dipping tempera-
scanning electron microscope (SEM), and the chemical compositions ture could be attributed to the Mn oxide film formed on the coating sur-
of the crystal phases in the coating was investigated using an energy- face during the hot dipping process. The presence of Mn oxides is
dispersive X-ray spectrometer (EDS). The phase of the coating surface evidenced by XPS/EDS/XRD characterization and more details will be
was analyzed by X-ray diffraction (XRD) using a Rigaku D/Max-2500V discussed in the following section. The formation of Mn oxide film
diffractometer with Cu Kα radiation and the monochromator operated would introduce a visible light interference effect with which a specific
at 40 kV and 200 mA. The scanning rate was set at 10°/min. The Mate- surface color will appear at certain film thickness and refractive index of
rials Data Inc. software Jade 5.0 and Powder Diffraction File were used the oxides. M. Tomita also reported that the color variations were due to
for phase identification. X-ray photoelectron spectroscopy (XPS) analy-
ses were also carried out using ESCALAB 250 (Thermo VG) with a Al Kα
source at 1486.6 eV in order to identify the film composition on the
coating surface.
The corrosion behavior of the surface and the ζ layers of the Zn coat-
ings with different Mn contents prepared at 450 °C/180 s were charac-
terized using an electrochemical workstation (Gamry Reference 600).
The hot dipped specimens were divided into two groups: as-dipped
specimens with no further treatment for the study of the coating's sur-
face corrosion resistance, and ζ layer-exposed specimens which were
obtained by fully removing the η layer from the as-dipped specimens
with #2000 silicon carbide grinding papers down to the ζ layer for the

Table 1
Chemical compositions of the prepared Zn alloys for hot dipping analyzed by XRF.

Alloy Zn Mn

Nominal Experimental Nominal Experimental

Zn 100 100 0 0
Zn–0.2Mn 99.8 99.785 0.2 0.215
Fig. 1. The coating colors of the steel work pieces hot dipped into (a) Zn at 450 °C/180 s,
Zn–0.5Mn 99.5 99.506 0.5 0.494
(b) Zn–0.5Mn at 430 °C/180 s, (c) Zn–0.5Mn at 450 °C/180 s and (d) Zn–0.5Mn at
Zn–0.7Mn 99.3 99.272 0.7 0.728
470 °C for 180 s, respectively.
Y. Wang, J. Zeng / Surface & Coatings Technology 245 (2014) 55–65 57

Fig. 2. (a) XPS survey spectra and high resolution spectra of (b) O 1s, (c) Zn 2p, and (d) Mn 2p for the pure Zn and Zn–0.5Mn coatings at different dipping conditions.

the different thicknesses of the Mn oxides film on the coating surface if dipping at different temperatures. As a result, the Mn oxide film
[24]. This is quite true in our case, since different times are needed for formed on the coating surface would have various thicknesses, and pos-
the specimens to be cooled till no further significant oxidation occurs sibly different oxidation levels, compositions of Mn oxides as well. These

(a) (b)

steel δ ζ η steel

25 µm 25 µm

(c) (d)

Fig. 3. SEM micrographs of the cross-section of hot-dip (a) Zn coating and (b) Zn–0.5Mn coating; (c) EDS line scan along the red line as shown in (b); (d) the magnified EDS line profile
corresponding to the dot-line highlighted area in (c).
58 Y. Wang, J. Zeng / Surface & Coatings Technology 245 (2014) 55–65

Table 2 subtle peak variations, as shown in Fig. 2(d), which may indicate the
EDS elemental compositions (wt.%) of the hot-dip Zn–0.5Mn coating at the positions indi- compositional difference of the Mn oxides in all these samples. The re-
cated in Fig. 3.
fractive index of the Mn oxide film would then be slightly different. At
Spot Zn Mn Fe this moment, we are not able to distinguish the oxide film thicknesses
A 89.27 0 10.73 of the specimens which may have only a few hundred nanometer vari-
B 91.84 0.42 7.74 ation according to the wavelengths of the light spectrum. But as de-
C 93.46 2.24 4.30 scribed previously, the different allowable oxidation times would
D 93.96 2.03 4.01
most likely lead to various oxide film thicknesses. According to the vis-
E 93.63 1.45 4.92
ible light interference effect theory, only the light having characteristic
wavelength determined by both the oxide film thickness and refractive
factors combined together would eventually lead to the distinct coating index could be observed. Therefore, the surface color of the specimen
surface colors of the specimens dipped at different temperatures. varies with dipping temperature.
XPS surface analyses of the Zn and Zn–0.5Mn coatings were carried The cross-section SEM micrographs of Zn and Zn–0.5Mn coatings
out trying to identify the compounds of the oxides on the coating. prepared at the hot-dipping conditions of 450 °C/180 s are shown
Fig. 2(a) shows the XPS survey spectra for the pure Zn and Zn–0.5Mn in Fig. 3(a) and (b). Layers with different Zn–Fe phases (δ-FeZn10,
coatings at different dipping conditions. It was found that the main ζ-FeZn13 and η-Zn) can be clearly distinguished on the hot-dip pure
peaks of Zn–0.5Mn coating in the XPS spectrum correspond to Zn, Mn, Zn coating, which is consistent with the previous report [1]. However,
O and C, respectively. Mn peak was not able to be detected in the pure the Γ phase mentioned in the report [1] is hardly to be observed in
Zn coating, and the other peaks of the Zn coating were similar in all Zn Fig. 3(a). It is most likely that the dipping time was too short for the
coatings with and without Mn addition. The intensities of Mn peaks Γ phase to be formed by diffusion [25]. Mn addition seems to have sig-
are far less than those of Zn, which is consistent with the fact of small nificant effects on both microstructure and Fe–Zn phase distribution
addition of Mn to the Zn coatings. High resolution XPS spectra of some in each layer of the coating. For example, the ζ layer of Zn–0.5Mn coat-
characteristic peaks would be necessary to differentiate the differences ing mainly consists of discontinuous small-sized ζ phases, as shown in
in these Zn coatings. Fig. 3(b), while the ζ layer of the pure Zn coating is in continuous
The high resolution XPS spectra of O 1s, Zn 2p, and Mn 2p are shown bulky columnar form (Fig. 3(a)). The morphological change of the
in Fig. 2(b–d). The C 1s peak at 284.6 eV corresponds to binding energy ζ layer is strongly associated with the Mn addition in the coatings. To
of the contaminated C. The O 1s peak at 531.3 eV suggests that the pres- better explain this, one might need to take a close look at the elemental
ence of metal oxides (Zn\O or Mn\O bonds). The Zn 2p3/2 peak at analysis of the coating. Table 2 lists the elemental compositions of
1022.1 eV is coincident with the literature value of the ZnO [22]. The five positions A, B, C, D and E indicated in Fig. 3(b). The results clearly
Mn 2p3/2 peak at around 641.5 eV is ascribed to Mn oxides [23]. Howev- show the presence of Mn at positions C, D and E in the ζ layer, and the
er, it was too difficult to deconvolve the Mn 2p peaks to identify the in- Mn contents (above 1.4 wt.%) are higher than 0.5 wt.% in the Zn–
dividual compositions of the Mn oxides due to very close binding 0.5Mn alloy bath for dipping. On the other hand, the Fe contents
energies of Mn for all the possible Mn oxides, such as MnO, Mn2O3, (in the range of 4–5 wt.%) at these positions are lower than the theoret-
and MnO2. However, careful inspection of the Mn 2p peaks of the spec- ical value of 6.2 wt.% in the pure FeZn13 phase, while the corresponding
imens prepared from different dipping temperatures would still give mass percent of Mn and Fe in the phase is about 6.3% in total and the

(a) (b)

200 µm 200 µm

(c) (d)

200 µm 200 µm

Fig. 4. SEM micrographs of the surface of as-dipped (a) pure Zn coating and (b) Zn–0.5Mn coating; (c) Mn and (d) O mapping of the Zn–0.5Mn coating shown in (b).
Y. Wang, J. Zeng / Surface & Coatings Technology 245 (2014) 55–65 59

Fig. 5. XRD patterns of the surface of as-dipped (a) pure Zn coating and (b) Zn–0.5Mn coating.

Zn contents of the position compositions are close to the theoretical formation of δ-FeZn10 phase in the liquid Zn phase. As a result, the thick-
value of 93.8 wt.% in the FeZn13 phase. Due to the fact of spot based ness of the δ layer was then increased in the Mn added Zn coatings.
EDS elemental analysis, multiple points other than the presented posi- The SEM micrographs of the as-dipped Zn and Zn–0.5Mn coating
tions in the ζ layer have been analyzed and the results were quite surfaces are shown in Fig. 4. As we can see from the figures, Mn addition
repeatable. All of these indicate partial replacement of Fe with Mn in has a significant effect on the microstructure of the coating surface. The
the FeZn13 phase of the Zn–0.5Mn coating. According to the Zn–Fe– grain boundary can be obviously observed on the Zn coating surface, but
Mn ternary phase diagram at 450 °C studied by Reumont et al. [26] not clearly for the Zn–0.5Mn coating. Fig. 4(c) and (d) shows the EDS
and Raghavan et al. [27], ζ-FeZn13 extends widely from the Fe–Zn binary mapping of the Mn and O elements on the Zn–0.5Mn coating, which
system to the Mn–Zn side because Mn could replace up to 75% of the indicates the uniform distribution of the Mn element throughout the
Fe in ζ-FeZn13 phase. Such replacement of Fe with Mn in the ζ phase surface. The content of Mn in the Zn–0.5Mn coating surface is around
could be accounted for the phase change from the continuous columnar 8.4 wt.% measured by the EDS mapping, much higher than the Mn con-
FeZn13 to discontinuous phase. tent of the as-prepared Zn–0.5Mn alloy. The result is consistent with the
To illustrate the Mn distribution at different layers of the Zn–0.5Mn previous test by EDS line scanning (see Fig. 3).
coating, EDS line scanning analysis along the red line shown in Fig. 3(b) The XRD patterns for the as-dipped pure Zn and Zn–0.5Mn coating
was performed and the corresponding data is presented in Fig. 3(c). surfaces are shown in Fig. 5. Indexing of the XRD pattern indicates
It could be confirmed that the Zn–0.5Mn coating layer consists of that the Zn coating contains Zn, ZnO and Zn–Fe phases, and the Zn–
three sub-layers from the surface to matrix steel, i.e., η, ζ and δ in 0.5Mn coating contains Mn, Zn, and Zn–Fe phases and Mn oxides
terms of Fe content. As we can see, Mn dominantly distributes in the ζ (for example, MnO). The Mn oxides formed on the Zn–0.5Mn coating
layer and top surface of the coating. The dot-line highlighted block surface is responsible for tuning the surface color of the coating (see
area in Fig. 3(c) with a spiking Mn content peak at the top surface is Fig. 1), as discussed previously. Zn–Fe phases are present in both pure
magnified in Fig. 3(d). This significantly increased Mn content suggests Zn and Zn–0.5Mn coatings, which can be explained that the iron diffuses
that Mn is segregated and enriched on the top surface of the η-layer and from the steel matrix to the molten Zn bath and forms Fe–Zn phases,
forms a Mn-rich sub-layer during hot dipping process. The thickness of particularly FeZn13 phase, in the molten Zn, and the FeZn13 phase ad-
the Mn-rich sub-layer is estimated at about 3 μm as shown by the line heres to the coating surface.
scan in Fig. 3(d). The formation of Mn oxides at the surface of Zn–0.5Mn coatings
Interestingly, the Mn distribution line in Fig. 3(c) shows that the could be related to the unevenly distributed Mn in the Zn melt. Mn is
Mn in δ-layer is lower than that in ζ-layer while the δ layer thickness a surface active element in the molten Zn, which tends to migrate to
(25 μm) of Zn–0.5Mn coating is larger than that (10 μm) of pure Zn coat- the surface from the bulk to form a Mn-rich layer at the surface of the
ing. Together considering the fact of increased thickness of δ layer with molten Zn. The Mn content at the surface of molten Zn could be more
Mn addition to the Zn coating, it indicates that Mn addition can promote than 10 times higher than that inside the melt and will be oxidized pref-
the growth of δ-FeZn10 phase. Similar phenomenon had been observed erentially due to the direct air exposure [29,30]. Hence, a Mn oxide film
in a previous work by N. Gao et al. [28]. They explained that the Mn ad- could be easily formed on the surface of the coating when withdrawing
dition in the zinc bath can stabilize the δ-FeZn10 phase and promote the the steel from the molten Zn–0.5Mn alloy at the high temperature.

(a) (b)
η-Zn
η-Zn

ζ-FeZn13

ζ-FeZn13

10 µm 10 µm

Fig. 6. SEM micrographs of the ζ layer-exposed (a) pure Zn coating and (b) Zn–0.5Mn coating.
60 Y. Wang, J. Zeng / Surface & Coatings Technology 245 (2014) 55–65

Fig. 7. (a) XPS survey spectra and (b) the high resolution Mn 2p spectra for the ζ-layer of the Zn coatings.

In order to observe the microstructure of the ζ layer of the coating, (Fe,Mn)Zn13 phase which has the non-distiguishable XRD pattern as
the coating was polished down to the ζ layer by means as described the FeZn13 crystal, and the low contents of Mn oxides and Mn contained
in the Experimental section. Fig. 6 shows the corresponding ζ-layer alloys which are below the XRD detection limit, if they do exist in the
microstructures of Zn and Zn–0.5Mn coatings. It appears that the ζ- ζ layer of Zn–0.5Mn coating.
layer of the Mn-free Zn coating still contains discrete Zn phase which
is surrounded by bulky FeZn13 phases. The ζ-layer of the Zn–0.5Mn 3.2. The corrosion properties
coating also contains FeZn13 phase and Zn phase, but the sizes of the
FeZn13 phase are much smaller than the FeZn13 phase in pure Zn coat- The Tafel polarization curves of the as-dipped and the ζ layer-
ing. This is consistent with the cross-section observation of the coating exposed Zn coatings with different Mn additions in 3.5% NaCl solution
in Fig. 3. The EDS analysis determined that the Fe content is around are shown in Fig. 9. The corresponding polarization parameters such
6.8 wt.% in the FeZn13 phase of the pure Zn coating which is very close as corrosion potential (Ecorr), corrosion current (Icorr), anodic polariza-
to the theoretical content of 6.2 wt.%. The Fe and Mn contents were tion slope (βa) and cathodic polarization slope (βc) for the as-dipped
estimated by the EDS analysis at around 4.4 wt.% and 2.1 wt.% in the and the ζ layer-exposed Zn coatings are shown in Tables 3 and 4,
FeZn13 phase of the Zn–0.5Mn coating. This is in line with the previous respectively.
elemental analysis of the FeZn13 phase in the cross-sectional microanal- The data in Table 3 reveal that the corrosion potentials (from
ysis (see Table 2). Same as previous discussion, partial Fe sites in FeZn13 − 0.996 to − 0.981 mV) of the as-dipped Zn coatings with different
phase were substituted by Mn to form ζ-(Fe,Mn)Zn13 phase. Mn contents are significantly higher than that (− 1.039 mV) of the
Fig. 7(a) shows the XPS survey spectra for the ζ layers of the pure Zn pure Zn coating. This seems to be opposite to the typical findings in elec-
and Zn–0.5Mn coatings. Similar to the surface of the as-dipped Zn coat- trodeposition obtained Zn–Mn coatings where the corrosion potential
ings, the Mn peaks can only be observed in ζ layer of Zn–Mn coating, became more negative with Mn addition since Mn is electrically a
and all other peaks are like for both Zn and Zn–Mn coatings, including more negative element than Zn [15,31]. However, in our case, all the
Fe peaks which are not present in the surface of all as-dipped Zn coat- as-dipped samples were not further polished to remove the oxide film
ings. The high resolution XPS spectra of Mn 2p are shown in Fig. 7(b). on the coating surface which actually provided an additional protection
The Mn 2p3/2 peak has the binding energy at around 640.7 eV, slightly layer to the as-dipped specimen. Mn oxides may provide higher passiv-
shifted from higher binding energy of Mn oxides, which may be as- ating potential which therefore lifts the apparent corrosion potential
cribed to Mn metal [23]. The XRD patterns of the ζ layers of Zn and for those Zn–Mn coatings. This could also contribute to the slightly
Zn–0.5Mn coatings are shown in Fig. 8. Indexing of the XRD pattern in- increased corrosion potential for the Zn coating with higher Mn addi-
dicates that the ζ layer of Zn and Zn–0.5Mn coatings mainly contain tion, as shown in Table 3, due to higher content of Mn oxides and film
FeZn13 and Zn phases. However, no Mn-containing phases could be thickness. As for the corrosion current of the as-dipped Zn coating, it
distinguished in the XRD patterns, while the EDS analysis in Fig. 3 decreases with the increasing of Mn addition. The corrosion current of
shows the existence of Mn. This is most likely due to the formation of the as-dipped pure Zn coating is 2.32 × 10−5 A·cm−2 which is much

Fig. 8. XRD patterns of the ζ layer-exposed (a) pure Zn coating and (b) Zn–0.5Mn coating.
Y. Wang, J. Zeng / Surface & Coatings Technology 245 (2014) 55–65 61

Fig. 9. Tafel polarization curves of (a) the as-dipped and (b) ζ layer-exposed Zn coatings with different Mn additions in 3.5% NaCl solution.

higher than that of Zn–Mn coating (around 3.0 × 10−6 A·cm−2 despite modify its properties [21,32–34]. As a result, the Zn–Mn coatings had
the Mn levels). This to some extent proves that the Mn addition can sig- lower cathodic polarization slope than that of pure Zn coating.
nificantly inhibit the anodic dissolution of Zn. Comparing the Tafel On the other hand, the polarization data of the ζ layer of the coatings
slopes of the as-dipped Zn–Mn coatings shown in Table 3, one can easily with different Mn additions listed in Table 4 indicate no significant
conclude that Mn addition decreases the cathodic polarization slope of change of the corrosion current for the ζ layers and the corresponding
the as-dipped coating, while it has no significant effect on the anodic po- magnitudes for the Mn-added Zn coatings were increased significantly
larization slope. To explain the slope tendencies with Mn addition upon partially due to loss of the passivating layer protection of oxides. As
cathodic and anodic polarizations, one must understand the involved for the corrosion potential of the ζ layer, it also shows the increase
corrosion reactions first. with the Mn addition, but with minor difference. This may indicate
In dilute solutions of salts such as NaCl, the following anodic and ca- slight enhancement of the corrosion resistance for the Zn–Mn coating.
thodic reactions would take place during the polarization process, and In fact, such increase may be resulted from the oxides on the surface
ZnO, Zn(OH)2, MnO and Mn(OH)2 are the common corrosion products of ζ layer which was formed during the η layer removal process. How-
within which the oxides have been observed to be the primary ones ever the quantity of the oxides would be much lower so that slight effect
[22]. has been introduced. The possible existence of oxides also leads to the
relatively lower anodic current density in Zn–Mn coatings. Similar as
2þ −
Anode : Zn→Zn þ 2e ð1  aÞ the observation for the as-dipped coatings, the anodic polarization
slopes for the ζ layer exposed specimens are almost identical. From
2þ − these results, it could be concluded with caution that the Mn addi-
Mn→Mn þ 2e ð1  bÞ
tion would possibly provide further positive anodic protection to the
Cathode : O2 þ 2H2 O þ 4e →4OH
− −
ð2Þ ζ layer from corrosion.
Although the data in Table 4 shows that the cathodic polarization
2þ −
slope also decreases with Mn addition to the Zn coating, the variation
Zn þ 2OH →ZnðOHÞ2 ð3  aÞ is smaller and the value is larger compared to those for the as-dipped
coatings. In fact, the cathodic polarization curves for all four specimens
2þ − are quite overlapped with almost the same cathodic polarization behav-
Mn þ 2OH →MnðOHÞ2 ð3  bÞ
ior. From this point, no significant effect of Mn addition on the cathodic
performance could be concluded, partially because of the much lower
ZnðOHÞ2 →ZnO þ H2 O ð4  aÞ
Mn content in the ζ layer and direct exposure of Zn phase in the
ζ layer to the corrosion environment.
MnðOHÞ2 →MnO þ H2 O: ð4  bÞ The EIS spectra for the surfaces of Zn coatings with different contents
of Mn are shown in Fig. 10(a1–a3). In Fig. 10(a3), the maximum phase
The oxides and hydroxides formed on the coating surface actually angles of the Zn, Zn–0.2Mn, Zn–0.5Mn, and Zn–0.7Mn coatings are esti-
serve as a barrier protection for the coatings from further corrosion. Re- mated at −55°, −60°, −60°, and −65°, respectively; and in Fig. 10(a2),
gardless of the thickness, the oxide film layer will provide very close the high impedance values at low frequencies increase from 600, 1900,
protection against moderate range of anodic polarization. Therefore, 2600, to 3000 Ω·cm2 with the addition of more Mn content. The EIS
change of the corrosion rate will not be significantly different during spectra for the ζ layers are shown in Fig. 10(b1–b3). Different from
the anodic polarization whether or not the Zn coating contains Mn. In that evaluated for the surface layer, the high impedance value at low
contrast, the decrease of cathodic polarization slope could be due the frequencies for the ζ layer of the Zn coating seems not to change with
fact that ZnO is a semiconductor and doping, and Mn addition can Mn content and maintains around 500 Ω·cm2, which is aligned with

Table 3
The polarization data of the as-dipped Zn coatings with different Mn additions in 3.5% NaCl solution.

Coating Ecorr (V) Icorr (A·cm−2) Tafel slope βa (V/decade) Tafel slope βc (V/decade)

Zn −1.039 2.32 × 10−5 2.24 × 10−2 1.568


Zn–0.2Mn −0.996 3.79 × 10−6 2.14 × 10−2 0.401
Zn–0.5Mn −0.992 2.92 × 10−6 2.50 × 10−2 0.369
Zn–0.7Mn −0.981 2.85 × 10−6 2.61 × 10−2 0.466
62 Y. Wang, J. Zeng / Surface & Coatings Technology 245 (2014) 55–65

Table 4
The polarization data of the ζ layer-exposed Zn coatings with different Mn additions in 3.5% NaCl solution.

Coating Ecorr (V) Icorr (A·cm−2) Tafel slope βa (V/decade) Tafel slope βc (V/decade)

Zn −1.050 1.46 × 10−5 2.53 × 10−2 1.16


Zn–0.2Mn −1.021 1.21 × 10−5 2.42 × 10−2 1.12
Zn–0.5Mn −1.013 1.10 × 10−5 2.65 × 10−2 0.92
Zn–0.7Mn −1.006 1.12 × 10−5 2.21 × 10−2 0.75

the conclusion of almost constant corrosion current for the ζ layer in the corresponding EIS spectra. The value that is normally correlated with
above polarization tests. the corrosion rate is the polarization resistance (Rp = Rct + Rpr). The
The EIS spectra of the surface and ζ layer can be interpreted in terms polarization resistances of the as-dipped coatings for Zn, Zn–0.2Mn,
of the equivalent circuit as shown in Fig. 11. In this model, Rs corre- Zn–0.5Mn, Zn–0.7Mn are 697, 2129, 3013, and 4040 Ω·cm2, respective-
sponds to the resistance of the solution, Cdl to the double-layer ca- ly, which indicates that the polarization resistance of the as-dipped
pacitance, and Rct to the charge-transfer resistance. The parameters of coating increases with Mn addition. Similarly, the corresponding
Rpr and Cpr element are related to the resistance and capacity of the model fitted data for the ζ layer are listed in Table 6. As we can see
corrosion products. The impedance of a phase element is defined as from the data, there is no significant variation in Rct and Rpr with the
ZCPE = [C(jw)n]−1, where 0 ≤ n ≤ 1. ZCPE represents an ideal capacity increasing of Mn addition, where the polarization resistances (Rp) of
of an interface when n = 1, and an ideal resistance of an interface ζ layers are about 550 Ω·cm2. The EIS results indicate that Mn addition
when n = 0. For the corrosion impedance, the value of n is associated can significantly improve the corrosion resistance of the surface of the
with the coating and affected by many factors, such as surface heteroge- as-dipped Zn coating, but has a little effect on the corrosion resistance
neity, Mn oxide film, and so on. of the ζ layer in the coating.
Table 5 summarizes the resistance, capacitance and n values of High resolution XPS spectra of Zn 2p and Mn 2p for pure Zn and Zn–
the coating surface by fitting the model described in Fig. 11 to the 0.5Mn coatings after Tafel polarization corrosion are shown in Fig. 12. As

Fig. 10. EIS spectra of (a1–a3) the as-dipped and (b1–b3) ζ layer-exposed Zn coatings with different Mn additions in 3.5% NaCl solution.
Y. Wang, J. Zeng / Surface & Coatings Technology 245 (2014) 55–65 63

Fig. 11. Equivalent circuit model used to fit EIS spectra.

we can see from Fig. 12(a), all the Zn 2p peak shifted from 1022.1 eV could inhibit further pitting corrosion and therefore improve the corro-
(which is associated with ZnO) to roughly 1022.8 eV after the polariza- sion resistance of the coating.
tion tests. The binding energy of 1022.8 eV is belonging to Zn(OH)2 — a Similarly, slight enhancement of corrosion resistance of ζ layer by
corrosion product on the coating surface [22]. Similarly, the binding Mn addition can also be visually observed by the SEM corrosion micro-
energy of Mn 2p3/2 peak for the Zn–0.5Mn coating after corrosion test graphs of ζ layers for the Zn and Zn–0.5Mn coatings after the Tafel polar-
increases from 641.4 eV (ascribed to MnO) to 642.5 eV, as shown in ization tests, as shown in Fig. 14. Possibly because of the relatively
Fig. 12(b), which indicates the formation of Mn(OH)2 on the coating higher corrosive resistance of the ζ-FeZn13 phase, corrosion tends to
after corrosion [23]. be focusing on the Zn phase area so that deep cracks can be observed
SEM micrographs of the surfaces of as-dipped Zn and Zn–0.5Mn on the ζ layer of the pure Zn coating, as shown in Fig. 14(a) and (b).
coatings after Tafel polarization tests were used to inspect the morpho- But still corrosion cracks could also be found in the ζ-FeZn13 phase. On
logical changes, which are presented in Fig. 13. Significant differences the other hand, in Fig. 14(c) and (d), for the ζ layer of the Zn–0.5Mn
could be identified between the presented surface corrosion micro- coating, pitting corrosion in the Zn phase can be observed, but no cracks
graphs of the as-dipped Zn and Zn–0.5Mn coatings. Pitting corrosion are visible in the FeZn13 phase. Interestingly, dense corrosion products
can be observed anywhere in the pure Zn coating, while dense corrosion can also be found anywhere in the ζ phase of the Zn–0.5Mn coating.
products can be found on the Zn–0.5Mn coating with few pitting corro- As indicated by previous discussion, the formation of MnO would be
sion. The ZnO tends to form a porous film on the coating surface and accounting for prevention of the FeZn13 phase from more serious corro-
allows the anions and cations to transport through the pores. The corro- sion. Due to the fact of more electrochemically active for Mn over
sion of Zn in deep site of the pores would have higher chance to contin- Zn [22], reaction (1-b) occurs prior to reaction (1-a). The preferential
ue and therefore results in the so-called pitting corrosion in the pure Zn dissolution of Mn can provide an active protection layer of MnO to
coating. But for the coatings with Mn addition, the MnO formed on the avoid further anodic dissolution of Zn. Besides, because of the much
surface tends to be more compact and dense, and leaves very few local higher solubility product constants (Ksp) for Mn(OH)2 [35], hydrolysis
pores. Therefore much less pitting corrosion could be observed on the of Mn2+ can be more difficult compared with Zn2+, which can reduce
as-dipped Zn–0.5Mn coating surface after polarization tests. It seems the local acidification inside the pores and therefore further reduce
that the dense corrosion products formed on the Zn–0.5Mn coating the anodic corrosion to prevent the pitting corrosion in the ζ-FeZn13

Table 5
The model fitted equivalent circuit components from the EIS spectra of the as-dipped Zn coatings with different Mn additions in 3.5% NaCl solution.

Coating Rs (Ω·cm2) Rpr (Ω·cm2) Cpr (Ω−1·cm−2·s−n) Rct (Ω·cm2) Cdl (Ω−1·cm−2·s−n)

CPE n1 CPE n2

Zn 7.4 471 2.5 × 10−5 0.83 226 2.8 × 10−3 0.95


Zn–0.2Mn 6.5 1414 2.1 × 10−6 0.76 715 0.9 × 10−3 0.87
Zn–0.5Mn 4.0 2283 3.9 × 10−6 0.70 730 1.2 × 10−3 0.94
Zn–0.7Mn 4.4 2260 3.1 × 10−6 0.75 1780 0.6 × 10−3 0.96

Table 6
The model fitted equivalent circuit components from the EIS spectra of the ζ layer-exposed Zn coatings with different Mn additions in 3.5% NaCl solution.

Coating Rs (Ω·cm2) Rpr (Ω·cm2) Cpr (Ω−1·cm−2·s−n) Rct (Ω·cm2) Cdl (Ω−1·m−2·s−n)

CPE n1 CPE n2

Zn 8.4 361 3.0 × 10−5 0.85 170 3.0 × 10−3 0.93


Zn–0.2Mn 11.5 351 4.0 × 10−5 0.73 205 1.2 × 10−3 0.94
Zn–0.5Mn 12.0 367 3.1 × 10−5 0.83 180 1.9 × 10−3 0.95
Zn–0.7Mn 11.4 344 3.9 × 10−5 0.72 240 1.1 × 10−3 0.92
64 Y. Wang, J. Zeng / Surface & Coatings Technology 245 (2014) 55–65

phase [22]. But the overall corrosion resistance would not be significant-
ly improved for the ζ-FeZn13 layer of Zn–0.5Mn coating due to the
presence of Zn phase, as evidenced by the Tafel polarization and EIS
data. For this reason, the Mn addition may have slight effects on improv-
ing the corrosion resistance of the ζ layer of Zn coatings.

4. Conclusions

The effects of Mn addition on the microstructures and corrosion per-


formance of both surface and ζ layer of the hot-dip Zn coatings have
been carefully investigated. The Mn oxide film can be easily formed on
the Zn–Mn coating surface during the hot dipping process. The surface
color of the coating is found varying with the dipping temperatures
and being attributed to the optical interference of the Mn oxide film.
The Mn distribution in δ layer is lower than that in the δ layer of the
hot-dip Zn–0.5Mn coating. Mn addition could lead to the increasing of
the δ layer thickness of the Zn coating. It is also found that Mn would re-
place partial Fe in the FeZn13 phase and change the FeZn13 phase from
the continuous columnar to discontinuous phase.
Both Tafel polarization and EIS tests indicate that Mn addition in the
coating can enhance the corrosion resistance of the as-dipped coatings.
The corrosion micrographs of the coating surfaces after the Tafel polar-
ization tests reveal the dense corrosion products formed on the surface
of the Zn–Mn coatings. It is assumed that the dense corrosion products
could inhibit the further pitting corrosion of the coating.
On the other hand, the corrosion potential of the ζ layer of the Zn
coating slightly increases with Mn addition, but the Mn addition has a
very limited effect on the corrosion current and impedance. As for the
ζ layers of both Zn and Zn–Mn coatings, the corrosion of Zn phase
Fig. 12. The high resolution XPS spectra of (a) Zn 2p and (b) Mn 2p for the coatings after seems to occur prior to the ζ phase. Dense corrosion products could
Tafel polarization tests. also be observed in the ζ phase of the coatings with Mn addition from
the SEM micrographs of the samples after polarization tests. To some
extent, Mn addition may help a little in the corrosion resistance of
ζ layer of the coating, but more investigation is needed.

(a) (b)

200 µm 10 µm

(c) (d)

200 µm 10 µm

Fig. 13. SEM micrographs of the surface of as-dipped (a, b) pure Zn coatings and (c, d) Zn–0.5Mn coatings after Tafel polarization tests; (b) and (d) are the magnified micrographs of se-
lected areas indicated by the red boxes in (a) and (c).
Y. Wang, J. Zeng / Surface & Coatings Technology 245 (2014) 55–65 65

(a) (b)

ζ-FeZn13

20 µm 5 µm

(c) (d)

ζ-FeZn13

20 µm 5 µm

Fig. 14. SEM micrographs of the ζ layer of hot-dip (a, b) pure Zn coating and (c, d) Zn–0.5Mn coating after Tafel polarization tests; (b) and (d) are the magnified micrographs of selected
areas indicated by the red boxes in (a) and (c).

Acknowledgment [16] N. Boshkov, K. Petrov, S. Vitkova, G. Raichevsky, Surf. Coat. Technol. 194 (2005)
276–282.
[17] P. Díaz-Arista, Z.I. Ortiz, H. Ruiz, R. Ortega, Y. Meas, G. Trejo, Surf. Coat. Technol. 203
This work is financially supported by the Guangxi Natural Science (2009) 1167–1175.
Foundation with the granted number of 2011GXNSFF018001. Prof. [18] C. Müller, M. Sarret, T. Andreu, Electrochim. Acta 48 (2003) 2397–2404.
[19] D. Sylla, C. Savall, M. Gadouleau, C. Rebere, J. Creus, P. Refait, Surf. Coat. Technol. 200
Hongwei Liu and Xiyong Chen at the Guangxi University are very (2005) 2137–2145.
much appreciated for constructive discussion and advices to this work. [20] N. Pistofidis, G. Vourlias, S. Konidaris, E. Pavlidou, A. Stergiou, G. Stergioudis, Mater.
Lett. 60 (2006) 786–789.
[21] L.D. Ballote, R. Ramanauskas, P. Bartolo-Perez, Corros. Rev. 18 (2000) 41–52.
References [22] B. Zhang, H.B. Zhou, E.H. Han, W. Ke, Electrochim. Acta 54 (2009) 6598–6608.
[23] M.C. Biesinger, B.P. Payne, A.P. Grosvenor, L.W.M. Lau, A.R. Gerson, Appl. Surf. Sci.
[1] A.R. Marder, Prog. Mater. Sci. 45 (2000) 191–271. 257 (2011) 2717–2730.
[2] X.G. Zhang, Corrosion and Electrochemistry of Zinc, Plenum Press, New York, 1996. [24] M. Tomita, S. Yamamoto, C. Tominaga, US Patent 5, 160, 552, (1992).
[3] C.E. Jordan, R. Zuhr, A.R. Marder, Metall. Mater. Trans. A 28 (1997) 2695–2703. [25] M. Onishi, Y. Wakamatsu, H. Miura, Trans. Jpn. Inst. Met. 15 (1974) 331–337.
[4] Y. Adachi, M. Arai, Mater. Sci. Eng. A 254 (1998) 305–310. [26] G. Reumont, G. Dupont, P. Perrot, Z. Metallkd. 86 (1995) 608–613.
[5] C.E. Jordan, A.R. Marder, J. Mater. Sci. 32 (1997) 5593–5602. [27] V. Raghavan, J. Phase Equilib. 24 (2003) 556–557.
[6] T. Kato, K. Nunome, K. Kaneko, H. Saka, Acta Mater. 48 (2000) 2257–2262. [28] N. Gao, Y.H. Liu, N.Y. Tang, J. Phase Equilib. Diffus. 31 (2010) 523–531.
[7] J.H. Selverian, A.R. Marder, M.R. Notis, Metall. Trans. A 19 (1988) 1193–1203. [29] C. Wagner, J. Electrochem. Soc. 99 (1952) 369–380.
[8] D. Phelan, B.J. Xu, R. Dippenaar, Mater. Sci. Eng. A 420 (2006) 144–149. [30] C. Wagner, J. Electrochem. Soc. 103 (1956) 627–633.
[9] J.H. Selverian, A.R. Marder, M.R. Notis, Metall. Trans. A 20 (1989) 543–555. [31] M. Bučko, J. Rogan, S.I. Stevanović, S. Stanković, J.B. Bajat, Surf. Coat. Technol. 228
[10] N. Pistofidis, G. Vourlias, S. Konidaris, E. Pavlidou, G. Stergioudis, Mater. Lett. 61 (2013) 221–228.
(2007) 2007–2010. [32] H. Leidheiser, I. Suzuki, J. Electrochem. Soc. 128 (1981) 242–249.
[11] J.D. Culcasi, P.R. Sere, C.I. Elsner, A.R. Di Sarli, Surf. Coat. Technol. 122 (1999) 21–23. [33] J. Vilche, K. Jüttner, W. Lorenz, W. Kautek, W. Paatsch, M. Dean, U. Stimming,
[12] N.C. Hosking, M.A. Ström, P.H. Shipway, C.D. Rudd, Corros. Sci. 49 (2007) 3669–3695. J. Electrochem. Soc. 136 (1989) 3773–3779.
[13] N. Boshkov, K. Petrov, S. Vitkova, S. Nemska, G. Raichevsky, Surf. Coat. Technol. 157 [34] P. Sharma, A. Gupta, K. Rao, F.J. Owens, R. Sharma, R. Ahuja, J.O. Guillen, B. Johansson,
(2002) 171–178. G. Gehring, Nat. Mater. 2 (2003) 673–677.
[14] S. Chang, J.C. Shin, Corros. Sci. 36 (1994) 1425–1436. [35] D.R. Lide, CRC Handbook of Chemistry and Physics, 84th ed. CRC Press, Boca Raton,
[15] N. Boshkov, Surf. Coat. Technol. 172 (2003) 217–226. 2003.

You might also like