Sunteți pe pagina 1din 12

Influence of Laser Processing Parameters on Microstructure

and Corrosion Kinetics of Laser-Treated ZE41 Magnesium Alloy


P. CHAKRABORTY BANERJEE, R.K. SINGH RAMAN, Y. DURANDET,
and G. McADAM

In the present study, surface melting of a magnesium alloy, ZE41, was performed with an
Nd:YAG laser using different laser parameters. The microstructure of the laser-treated and
untreated specimens was analyzed by optical and scanning electron microscopy and X-ray
diffraction. Corrosion resistance of the different laser-treated specimens along with the
untreated alloy was characterized using electrochemical impedance spectroscopy and weight loss
measurements in 0.001 M sodium chloride solution. Although the laser processing parameters
influenced the microstructure and the melt depth of the laser-treated zone, these had little effect
on the corrosion resistance of the alloy.

DOI: 10.1007/s11661-012-1590-x
 The Minerals, Metals & Materials Society and ASM International 2013

I. INTRODUCTION trolled quantum of energy or power with a precise


spatial distribution, short processing time, flexibility of
THE attractive mechanical properties of Mg and its operation, economy of material consumption, forma-
alloys have prompted their increasing use in automobile tion of a relatively narrow heat-affected zone, and
and aerospace applications. However, their poor corro- precision of operation.[18,29]
sion resistance has restricted their wider use. Alloy Laser surface treatments have been reported to
microstructure has a profound influence on the corro- improve the corrosion and pitting resistance of various
sion properties of any alloy system. Generally, the alloy systems[18–26,32–37] as well as Mg alloys.[18–26]
a-solid solutions of Mg alloys are anodic to the interme- Majumdar et al.[18] reported an improvement in pitting
tallic particles and are preferentially corroded,[1–9] resistance of MEZ alloy after laser surface melting and
except in the case of Mg-8Li alloy,[10] where the interme- attributed this improvement to the combined effects
tallic particles are more anodic than the a-solid solution. of grain refinement, visible absence of reprecipitation of
Suitable tailoring of the surface microstructure and/or the cathodic intermetallic phases, and enrichment of
chemistry is a common approach to improving the alloying elements in the solid solution. Coy et al.,[19]
corrosion resistance of a susceptible metal/alloy. For this Guan et al.,[20] and Gao et al.[21] have reported
purpose, different rapid solidification processes have been improvements in corrosion resistance of AZ91D alloy
employed for various Mg alloys.[11–30] In contrast to the as a result of laser surface melting. However, Guan
conventional casting processes, rapid solidification pro- et al.[20] and Gao et al.[21] reported a difference in the
cesses generate a highly refined microstructure with a fine corrosion resistance of the overlapping and the non-
dispersion of intermetallic particles and produce a overlapping zones of the laser tracks. Guo et al.[22]
homogeneous distribution of alloying elements with attributed the improvement in polarization resistance of
extended solid solutions.[31] the laser-melted WE43 alloy to the dissolution of the
Among the various rapid solidification processes, cathodic Mg12Nd particles during laser surface melting.
laser surface modifications (melting/alloying/cladding) Liu et al.[23] have also reported improvement in the
provide several advantages, such as delivery of con- corrosion resistance of a cerium (Ce)-containing AZ
series Mg alloy as a result of overall microstructural
refinement, increased aluminum (Al) concentration in
P. CHAKRABORTY BANERJEE, formerly Ph.D. Student with the melted layer, and formation of a Ce-containing
the CAST Cooperative Research Centre, Hawthorn, VIC 3122, corrosion-resistant surface film. Abbas et al.[24] and
Australia, and Department of Chemical Engineering, Monash Uni-
versity, Melbourne, VIC 3800, Australia, is now Post-doc Fellow with Singh Raman et al.[25] attributed the improved corrosion
the Department of Mechanical and Aerospace Engineering, resistance of AZ31, AZ61, AZ91, and WE43 alloys to
Monash University. R.K. SINGH RAMAN, Professor, is with the the pronounced grain refinement and redistribution of
Department of Chemical Engineering, Monash University, and also the intermetallic particles. Koutsomichalis et al.[26]
with the Department of Mechanical and Aerospace Engineering,
Monash University. Contact e-mail: raman.singh@monash.edu
observed similar improvement in the case of AZ31B
Y. DURANDET, Research Fellow, is with the Industrial Research alloy. However, Banerjee et al.[29] reported the improve-
Institute Swinburne (IRIS), Swinburne University of Technology, ment in corrosion resistance of the laser surface-melted
Hawthorn, VIC 3122, Australia. G. McADAM, Senior Research ZE41 alloy to be insignificant as compared to the Al-
Scientist, is with the Defence Science and Technology Organisation, containing Mg alloys and attributed this behavior to the
Fishermans Bend, VIC 3207, Australia.
Manuscript submitted February 27, 2012. absence of any beneficial alloying elements in the
Article published online January 12, 2013 former. In fact, Dubé et al.[27] reported little improvement

2346—VOLUME 44A, MAY 2013 METALLURGICAL AND MATERIALS TRANSACTIONS A


in the corrosion resistance even in the case of Al- II. EXPERIMENTAL PROCEDURES
containing alloys (AZ91D and AM60B), although
refinement of microstructure had been achieved. The A. Test Material
microstructure of the laser-melted surface depends Rectangular plates (93 9 65.5 9 6 mm3) of ZE41A
strongly on the laser processing parameters, viz., laser alloy were used in T5 condition. The nominal alloy
scan speed, power, etc.[18,27] In the study by Dubé composition as provided by Spectrometer Services Pty
et al.,[27] the microstructure produced for the two alloys Ltd is given in Table I.
with their laser parameters consisted of Al-rich curled
bands, which created galvanic micro-currents and there-
by enhanced the dissolution of their neighboring zones B. Laser Processing
within the melt pool. It is clear that the laser processing An Nd:YAG laser operating in a continuous wave
parameters alter the morphology of the secondary mode was used as the heat source. The sample was
phases and, in turn, the corrosion resistance of the Al- clamped on a stage, which was moved at a desired speed
containing Mg alloys. (to achieve the desired scan rate), while the laser was
Investigations of the role of laser surface melting on fixed. A gas cone with nozzle was used to protect the
corrosion resistance have largely been carried out on Al- focussing lens from dust and splatters and shroud the
containing alloys, and Al-free Mg alloys have received limited laser spot area to avoid oxidation during laser melting.
attention.[18,22,24,29] Durandet et al.[30] investigated the influ- Argon was used as the shielding gas at a flow rate of
ence of laser processing parameters on the microstructure of 3 l/min at the gas cone inlet. The laser beam was
ZE41 alloy, but did not report any corrosion data. focussed from a single spot to two twin spots by adding
In the present study, the influence of laser processing a dual spot lens to the system. The nominal power split
parameters (such as laser power, scan rate, and beam was 50:50 for the dual spot.
dimensions) on the resultant microstructure has been In the present study, to understand the influence of
investigated. The influence of the resultant microstruc- cooling rate and heat input in terms of laser power
ture on the corrosion resistance of the laser-treated density (laser power/area of the beam), or line energy
ZE41 alloy at different durations of immersion in (laser power/scan rate) on the resultant microstructure
0.001 M NaCl has been investigated. An electrical and hence on the corrosion behavior, the scan rate and
equivalent circuit (EEC) has been developed to obtain beam dimensions have been varied keeping the laser
a mechanistic understanding of the time-dependent powers similar. The laser parameters are given in
electrochemical processes at the different interfaces of Table II.
the laser-treated and untreated alloy.
Table I. Nominal Composition of ZE41 (in wt pct)

Mg Zn Ce La Nd Pr Zr Fe Cu Ni
Bal. 3.82 0.64 0.27 0.14 0.06 0.69 0.002 0.002 <0.001

Table II. Laser Processing Parameters

Specimen Nominal Laser Scan Rate Beam Dimensions Nominal Power Nominal Line
Specification Power (W) (m/min) Beam Profile (mm) Density (W/mm2) Energy (J/mm)
Sample 1 1856 19.875 D=1 1182 5.6

Sample 2 1818 0.5 a = 6.5 89 218.2


b=1

METALLURGICAL AND MATERIALS TRANSACTIONS A VOLUME 44A, MAY 2013—2347


C. Microstructure of the Laser-Treated
and Untreated ZE41
The microstructures of both the laser-treated and
untreated alloy were examined using optical microscopy
and scanning electron microscopy (SEM). SEM analyses
were carried out using a JEOL 840A scanning electron
microscope operating at 20 keV. The specimens were cold
mounted ‘‘edge on’’ using epoxy resin in order to polish
and etch both the treated and untreated areas of the cross
section. This facilitated a direct comparison between the
substrate (untreated zone) microstructure and the laser-
melted zone microstructure and enabled the measurement
of the depth of the laser-treated zone. The specimens were
etched in 78.6 pct ethylene glycol + 20.6 pct of water and
0.8 pct of HNO3 for 2 minutes. Corrosion morphologies of
the laser-treated and untreated specimens were examined
using a JEOL 840A SEM. The specimens were analyzed Fig. 1—Electrochemical cell with three electrodes (the specimen with
without cleaning the corrosion products (after the corro- an exposed area of 0.785 cm2 was used as the working electrode,
sion tests, the specimens were rinsed with deionised water whereas a platinum mesh and a saturated calomel electrode were
and were subsequently dried with compressed air). All the used as the counter and reference electrodes, respectively).
corroded specimens were gold coated prior to SEM to
avoid charging of the less conductive corrosion products. amplitude of 10 mV. The impedance response was
The phase composition of the untreated and laser- measured over frequencies between 1 and 10 mHz,
treated alloy was analyzed by XRD, using a Cu Ka recording 10 points per decade of frequency using PAR
radiation, with a scan over 20 to 60 deg (2h) at a scan PowerSuite Electrochemistry package. EIS tests were
rate of 2 deg/min and a step size of 0.02. performed at different times of immersion without any
intentional change in the pH of 0.001 M NaCl solution.
D. Corrosion Study of the Laser-Treated Impedance analysis was carried out using PAR ZSimp-
and Untreated ZE41 Win package for Windows for frequencies between
100 kHz and 100 mHz to prevent misinterpretation of
To perform electrochemical testing, untreated ZE41 any artifacts that are generally present in the high
specimens were ground using SiC papers (800–2500 frequency region as well as the scatter that occurs in the
grit), whereas only 2500 grit paper was used for the low frequency region. All the electrochemical tests were
laser-treated specimens in order to avoid the possible repeated at least thrice to examine the reproducibility of
grinding away of the relatively thin laser-treated zone results.
during coarser grinding. The ground specimens were In order to perform weight loss measurements, the
polished with 3 and 1 lm diamond suspensions, rinsed laser-treated and the untreated specimens were im-
with deionised water, degreased with acetone, and dried mersed in 0.001M NaCl for 7 days (168 hours). All the
with compressed air. In grinding/polishing the laser- specimens were cold mounted using epoxy resin. The
treated specimens, extra care was taken to insure that junction of the cold mounting and the alloy surface was
the laser-treated part was not ground/polished away. enamel coated to avoid any electrolyte penetration at
The melt depths of the laser-treated specimens were this junction. The corroded samples were cleaned with
measured before and after grinding and polishing. The chromic acid solution (200 g/L Cr2O3 + 20 m/l
grinding and polishing removed an average melt depth Ba(NO3)2 + 10 g/L AgNO3) according to the ASTM
of around 50 lm. Standard G1-90, Designation C5.2.[38] The corrosion
EIS tests were performed in 0.001 M NaCl solution rates were evaluated using the following formula[39]
using a Princeton Applied Research (PAR) potentiostat
(Model 2273) and an electrochemical cell with three 87:6 W
mm=y ¼
electrodes (the specimen with an exposed area of DAT
0.785 cm2 was used as the working electrode, whereas
where W is a weight loss in mg; D is the density of the
a platinum mesh and a saturated calomel electrode were
specimen in g/cm3; A is the exposed area of the specimen
used as the counter and reference electrodes, respec-
in cm2; and T is the time in hours.
tively). The experimental setup is shown in Figure 1. All
the experiments were performed at the ambient temper-
ature. Before every experiment, the open circuit
potential was monitored for 1 hour to insure a stable III. RESULTS AND DISCUSSION
electrochemical condition. A fluctuation of less than
A. Microstructure of the Laser-Treated
10 mV of the open circuit potential sustained for a
and Untreated ZE41
period of 1000 s was considered as the stable electro-
chemical condition. The EIS tests were carried out by The microstructure of the untreated alloy (Figure 2(a))
applying a sinusoidal potential wave at Ecorr with an consists of an a-solid solution and intermetallic particles

2348—VOLUME 44A, MAY 2013 METALLURGICAL AND MATERIALS TRANSACTIONS A


Fig. 2—Back scattered electron image of the top surface of ZE41 alloy: (a) without laser treatment, (b) faster scan rate specimen (sample 1), and
(c) slower scan rate specimen (sample 2) (Refer to Table II for laser parameters used to produce samples 1 and 2).

along the grain boundaries. The average grain size of the laser-treated zone of both the specimens, forming a
untreated alloy was approximately 50 lm. Figures 2(b) refined and continuous network of precipitates. Figure 3
and (c) show the microstructures of the laser-treated shows the XRD scans of the laser-treated and untreated
samples 1 and 2, respectively. Since the scan rate used for ZE41 alloy. It is evident that both the laser-treated
sample 1 was faster than that for sample 2, while the laser specimens had similar phases (i.e., a–Mg and
power was similar, the cooling rate would be more rapid Mg7Zn3(RE)) as that of the untreated alloy. The phase
in the case of the faster scan rate specimen (sample 1) and composition of the untreated alloy is in agreement with
hence the degree of microstructural refinement achieved earlier studies.[7,8]
was greater in the case of sample 1. On the other hand, a The cross sections of the two laser-treated specimens
more homogeneous microstructure was obtained in the are shown in Figure 4. Large surface cracks were
case of the slow scan rate specimen (sample 2) because of observed in the case of the faster scan rate specimen
the increased laser dwell time, which is given by[30] (sample 1), while the slower scan rate specimen (sample
2) was free of optical microscopically discernable cracks
L
t¼ ½1 (Figure 4(b)). Since ZE41 alloy has a high coefficient of
V thermal expansion and shrinkage due to solidifica-
where L is the beam interaction length in the direction of tion,[40] the presence of cracks in the faster scan rate
laser scan and V is the scan rate. The laser dwell time of specimen (sample 1) may be attributed to the greater
the slower scan rate specimen (2 9 103 min) was about power density and faster scan rate (Table II) employed,
an order of magnitude greater than the faster scan rate which provided a high input energy and a quicker
specimen (1 9 104 min). This led to an increased time cooling rate, causing a high thermal stress and cracking
in the liquid phase for dissolution of second phases and in the laser-treated zone. The melt depths of the slower
diffusion, and as a result a superior melt homogeniza- scan rate specimen (sample 2: ~750 lm) were consider-
tion occurred in the case of the slower scan rate ably greater than that of the faster scan rate specimen
specimen.[30] (sample 1: ~250 lm), which can be attributed to the
The intermetallic particles reprecipitated along the greater line energy and laser dwell time experienced at
grain boundaries of the refined structure throughout the the slower scan rate.

METALLURGICAL AND MATERIALS TRANSACTIONS A VOLUME 44A, MAY 2013—2349


faster scan rate specimen (sample 1)
slower scan rate specimen (sample 2)
Untreated ZE41

500
α - Mg

Mg7Zn3(RE)
Intensity (CPU) 400

300

200

100

0
20 25 30 35 40 45 50 55 60
2θ (degree)

Fig. 3—XRD of the laser-treated and untreated alloy.

B. Electrochemical Impedance Spectroscopy


of the Laser-Treated and Untreated Alloy
The impedance plots (Nyquist, Bode modulus and
phase angle) of the untreated and laser-treated alloy at
3 hours of immersion in 0.001 M NaCl are shown in
Figure 5.
As in the case of Mg and Mg alloys,[1,41,42] the
Nyquist plots for both laser-treated and untreated ZE41
are characterized by two capacitive semicircles. The high
frequency semicircles correspond to charge transfer
processes, and the medium frequency semicircles are
attributed to mass transfer through the corrosion
product layer. The corrosion resistance of any material
is determined by the combined diameters of the two
capacitive loops.[41] It is evident from Figures 5(a) and
(b) that the total impedance of the laser-treated speci-
mens was marginally higher than that of the untreated
specimens (the scatter in the data was ~2 kX cm2 for
both the Nyquist and the Bode modulus plots).
Figures 6 and 7 show the Nyquist plots and the Bode
(modulus and the phase angle) plots of the laser-treated
and untreated specimens at different durations of
immersion in 0.001 M NaCl. At 7 hours of immersion,
the impedance values (Figures 6 and 7) of both the laser-
treated specimens were 2 kX lower than those at 3 hours
of immersion (Figure 5), but were similar to that of the
untreated specimen. No significant change in the imped-
ance was observed for all the specimens immersed
for durations between 7 and 25 hours. However, at
55 hours of immersion, the impedance of each of them
was lower (~2 kX) than the corresponding impedance at
25 hours. While the specimen produced with a slower
Fig. 4—Optical micrographs of the cross section of ZE41 alloy laser
melted with different laser parameters (as described in Table II):
scan rate had similar corrosion resistance (scatter in the
(a) faster scan rate specimen (sample 1) and (b) slower scan rate data ~1.5 kX cm2 for both the Nyquist and the Bode
specimen (sample 2). modulus plots) as that of the untreated specimen, the

2350—VOLUME 44A, MAY 2013 METALLURGICAL AND MATERIALS TRANSACTIONS A


20 Untreated parallel with a resistance. This forms the basis of the
faster scan rate (sample 1) EEC model that is employed. An EEC was developed
slower scan rate (sample 2) based upon hypothetical corrosion mechanisms and
15

-Z" (kΩ cm2)


simulated to correspond to the experimental data.
Theoretical simulation was conducted using the EEC
10 as shown in Figure 8. In this EEC, Rs is the solution
resistance and the surface film is represented by a
5 parallel combination of the film capacitance, Cf, and the
film resistance, Rf. The surface film is in a parallel
0
combination with the electrical double layer. The
0 5 10 15 20 electrical double layer is represented by a constant
Z' (kΩ cm2) phase element (CPE), Qdl, and a charge transfer
(a) resistance, Rc. The electrochemical behavior of the
CPE is generally attributed to the distributed surface
30 reactivity, roughness, electrode porosity, and current
Untreated
faster scan rate (sample 1) and potential distributions associated with the electrode
25
slower scan rate (sample 2) geometry.[43] This model was successfully used earlier in
|Z| (kΩ cm2)

20 a separate study to simulate the impedance data of ZE41


alloy.[29]
15
The proposed EEC (Figure 8) was the closest fit to the
10 experimental impedance data for the untreated and all
the laser-treated specimens, with the lowest Chi square
5
values and the maximum errors in measurements being
0 less than 4 pct. The maximum errors in simulated data
0.1 1 10 100 1000 10000 100000 1000000
for the untreated and the laser-treated specimens were
Frequency (Hz) less than 7 to 9 pct in |Z| and 6 to 7 pct in angle. A
(b) representative error plot (error plot of the faster scan
rate specimen at 3 hours of immersion in 0.001 M NaCl)
80 is shown in Figure 9. Error plots for all other specimens
Untreated
Phase angle (degree)

70 faster scan rate (sample 1) at different times of immersion were similar to the
slower scan rate (sample 2)
60 representative plot.
50 It is evident from the representative comparisons in
40 Figure 10 that the simulated impedance plots (Bode
30 plots), that were generated using the EEC in Figure 8,
20
were in good agreement with the experimental plots
obtained at different times of immersion. Although the
10
impedance data for all the specimens were recorded in
0
0.1 1 10 100 1000 10000 100000 1000000 the frequency range 1 MHz to 10 mHz, the impedance
Frequency (Hz) data were generally simulated in the frequency range of
100 kHz to 100 mHz in order to avoid any artifact in the
(c) high frequency range and/or noise in the low frequency
range,[44] and this practice is consistent with several
Fig. 5—Impedance plots of the laser-treated and untreated speci- other researchers.[41,45–47] The interfacial capacitances
mens at 3 hours of immersion in 0.001 M NaCl: (a) Nyquist, (b)
Bode modulus, and (c) Bode phase angle plots. and resistances associated with the EEC (Figure 8) were
calculated in order to understand the electrochemical
kinetics at the different interfaces (the corrosion product
specimen produced with a faster scan rate had the lowest film interface and the electrical double layer).
corrosion resistance. This may be attributed to the Figure 11 shows the evolution of Cf and Rf of all the
presence of cracks in the case of this specimen, which specimens at different times of immersion. It is evident
was generated by a faster scan rate and higher power that Cf and Rf of all specimens were similar and
density (laser power/area of the beam). remained reasonably unaltered with time, suggesting
EIS is a powerful tool to investigate the nature of no significant change in the hydroxide layer/solution
surface films and associated corrosion mechanisms of interface at different durations of immersion. The
metals and alloys. In order to attain a mechanistic evolution of Qdl and Rc is shown in Figure 12. Qdl of
understanding of the influence of laser parameters on all the specimens increased with time. The exponent (n)
the time-dependent corrosion process, a thorough anal- value of Qdl was close to 1, so it can be approximated as
ysis of the impedance data was carried out. Hence, in the a capacitance.[43] Since capacitance is directly propor-
present study, the impedance of substrate (both un- tional to the area of the capacitor, the trend of increase
treated and laser-treated)/corrosion product film/NaCl of Qdl over time for both the specimens indicates an
solution interfaces was analyzed using electrical equiv- increasing area fraction of exposure of the metal/
alent circuits (EEC). In corrosion processes, each of the hydroxide interface to the electrolyte with time. Qdl of
two interfaces behaves as a capacitance combined in all the specimens was similar until 7 hours of immersion;

METALLURGICAL AND MATERIALS TRANSACTIONS A VOLUME 44A, MAY 2013—2351


20 faster scan rate (sample 1) 20 faster scan rate (sample 1)
slower scan rate (sample 2) At 7 h slower scan rate (sample 2) At 25 h
Untreated
15 15 Untreated
-Z" (kΩ cm2)

-Z" (kΩ cm2)


10 10

5 5

0 0
0 5 10 15 20 0 5 10 15 20
Z' (kΩ cm2) Z' (kΩ cm2)

20 faster scan rate (sample 1)


slower scan rate (sample 2) At 55 h
15 Untreated
-Z" (kΩ cm2)

10

0
0 5 10 15 20
Z' (kΩ cm2)

Fig. 6—Nyquist plots of the laser-treated and the untreated ZE41 at different times of immersion in 0.001 M NaCl.

however, the slower scan rate specimen (sample 2) had of the faster scan rate specimen was a little higher than
lower Qdl than the faster scan rate (sample 1) and the the other two specimens. The slightly higher corrosion
untreated specimens, suggesting a smaller area fraction rate observed in the case of the faster scan rate specimen
of exposure of the metal/hydroxide interface than the was attributed to the cracks on its surface. This trend in
other two specimens. Rc of the faster scan rate specimen the corrosion resistance of the laser-treated and un-
(sample 1) had a decreasing trend, whereas Rc of the treated specimens is consistent with that obtained using
slower scan rate specimen (sample 2) and the untreated the electrochemical method (EIS).
specimen remained constant with the increase in immer- Figure 15 shows the typical evolution of the electrolyte
sion time. pH with increasing immersion time. The change in
The polarization resistance (Rp) is represented as the electrolyte pH was similar for all the specimens. After
algebraic sum of the corrosion film resistance and the 1 day (24 hours) of immersion, the initial pH (pH 5) of the
charge transfer resistance.[41] Rp of all the specimens electrolyte increased to pH 9 and was constant for the rest
(Figure 13) was similar till 7 h; however, after 24 hours of the test duration. This suggests that during the initial
of immersion, the Rp of the faster scan specimen hours of immersion (1 day), there was local alkalization
(sample 1) was the lowest, whereas the Rp values of due to active metal dissolution and the formation of
the untreated ZE41 and the slower scan rate specimen hydroxide film on the metal surface. This is also supported
(sample 2) were similar. The lower Rp of the faster scan by the Rc value (Figure 12(b)), which suggests that the
rate specimen (sample 1) is attributed to the presence of charge transfer resistance decreased (due to active metal
cracks on its surface. Although a refined, homogeneous, dissolution) until 24 hours of immersion (1 day), but
and crack-free microstructure was obtained in the case remained constant after that (due to the formation of the
of the slower scan rate specimen (sample 2), the corrosion product film). Furthermore, the constancy of
polarization resistance of this specimen was similar to the electrolyte pH agrees well with the constancy in the
that of the untreated ZE41. values of the parameters (Cf and Rf) related to the
corrosion product film/electrolyte interface.
C. Weight Loss Measurements
A comparison of the corrosion rates of the laser- D. Corrosion Morphology
treated and untreated specimens is shown in Figure 14. Figure 16 shows the corrosion morphology of the
The corrosion rates of the slower scan rate specimen and untreated and the laser-treated specimens at 55 hours
the untreated specimen were similar after 7 days of immersion in 0.001 M NaCl. Thick corrosion prod-
(168 hours) of immersion, whereas the corrosion rate ucts with scattered pits covered the entire surface of

2352—VOLUME 44A, MAY 2013 METALLURGICAL AND MATERIALS TRANSACTIONS A


80 faster scan rate (Sample 1)
30 faster scan rate (Sample 1) At 7 h
At 7 h 70 slower scan rate (Sample 2)

Phase angle (degree)


slower scan rate (Sample 2) Untreated
25 Untreated 60

|Z| (kΩ cm2)


20 50
40
15
30
10
20
5 10

0 0
0.1 10 1000 100000 10000000 0.1 10 1000 100000 10000000

Frequency (Hz) Frequency (Hz)

30 80 faster scan rate (Sample 1)


faster scan rate (Sample 1) At 25 h
slower scan rate (Sample 2) At 25 h 70 slower scan rate (Sample 2)
25 Untreated Untreated

Phase angle (degree)


60
|Z| (kΩ cm2)

20
50
15 40

10 30

20
5
10
0
0.1 10 1000 100000 10000000 0
0.1 10 1000 100000 10000000
Frequency (Hz)
Frequency (Hz)
30 faster scan rate (Sample 1) 80 faster scan rate (Sample 1)
At 55 h At 55 h
slower scan rate (Sample 2) 70 slower scan rate (Sample 2)
25 Untreated Untreated
Phase angle (degree)

60
20
|Z| (kΩ cm2)

50
15 40

30
10
20
5
10

0 0
0.1 10 1000 100000 10000000 0.1 10 1000 100000 10000000
Frequency (Hz) Frequency (Hz)

Fig. 7—Bode impedance plots of the laser-treated and untreated specimens at different times of immersion in 0.001 M NaCl.

Cf the grains remained intact. However, numerous pits


were observed on the corrosion product film of this
specimen. Similar to the untreated alloy, a thick
Rs corrosion film with numerous small pits covered the
Q dl
entire surface of both the laser-treated specimens.
Laser surface melting has resulted in an improvement
Rf
in the corrosion resistance of several Mg alloys.[18–26,29]
Rc The improvements have been attributed to the changes
in morphological and microchemical features of the
alloy microstructure by significantly refining the surface
Fig. 8—The electrical equivalent circuit of the untreated and laser- microstructure,[18,20,21,23–26] enhancing the beneficial
treated ZE41 alloy in 0.001 M NaCl (Rs is the solution resistance, Cf alloying elements (such as aluminum) in the solid
is the film capacitance, Rf is the film resistance, Qdl is the CPE asso-
ciated with the electrical double layer, and Rc is the charge transfer
solution,[18–21,23,24] and dissolving[19,22] and/or redistrib-
resistance). uting the intermetallics in the laser-treated zone. A
preliminary study by Banerjee et al.[29] suggested an
each specimen. In the case of the untreated ZE41 insignificant improvement in corrosion resistance of
(Figure 16(a)), some of the grains were completely ZE41 alloy as a result of laser treatment. However, the
corroded (as shown by the arrows), whereas a few of study was carried out with a narrow set of laser

METALLURGICAL AND MATERIALS TRANSACTIONS A VOLUME 44A, MAY 2013—2353


Fig. 9—Error plots of the faster scan rate specimen (sample 1) at 3 h of immersion in 0.001 M NaCl.

Fig. 10—Comparisons of the Bode modulus and phase angle plots at different durations of immersion: experimentally determined and simulated
data for a typical laser-treated specimen (sample 1 in this case).

parameters. Even though the present study employed a confirms the earlier proposition that in the absence of
wide variation in laser parameters, the resulting micro- any beneficial alloying elements (such as Al) in the ZE41
structure (Figure 2) still produced no observable alloy, the influence of the change in microstructure
improvement in corrosion resistance. This further due to laser processing parameters on the corrosion

2354—VOLUME 44A, MAY 2013 METALLURGICAL AND MATERIALS TRANSACTIONS A


0.5 25
Faster scan rate (sample 1) Faster scan rate (sample 1)
Slower scan rate (sample 2) Slower scan rate (sample 2)
0.4 20
Untreated ZE41 Untreated ZE41

Cf (nF/cm2)

Rf (kΩ cm2)
0.3 15

0.2 10

0.1 5

0 0
0 10 20 30 40 50 60 0 10 20 30 40 50 60
Time (h) Time (h)

(a) (b)

Fig. 11—Evolution of (a) Cf and (b) Rf of the laser-treated and untreated specimens in 0.001 M NaCl solution.

250 20 Faster scan rate (sample 1)


Faster scan rate (sample 1)
Slower scan rate (sample 2) Slower scan rate (sample 2)
200 Untreated ZE41 Untreated ZE41
15
Qdl (µF/cm2)

Rc (kΩ cm2)
150
10
100

5
50

0 0
0 10 20 30 40 50 60 0 10 20 30 40 50 60

Time (h) Time (h)


(a) (b)

Fig. 12—Evolution of (a) Qdl and (b) Rc of the laser-treated and untreated specimens in 0.001 M NaCl solution.

Faster scan rate (sample 1) 1.6


Slower scan rate (sample 2) 1.4
Corrosion rate (mm/y)

20 Untreated ZE41
1.2

1
15
Rp (kΩ cm2)

0.8

10 0.6

0.4

5 0.2

0
0 Faster scan Slower scan Untreated
0 10 20 30 40 50 60 rate rate ZE41
Time (h)
Fig. 14—Weight loss measurements of the laser-treated (the faster
Fig. 13—Evolution of Rp of the laser-treated and untreated speci- scan rate and the slower scan rate) and untreated ZE41.
mens in 0.001 M NaCl solution.

resistance of this alloy is insignificant.[29] The thorough chemical parameters (Figures 11 and 12) among the
time-dependent EIS characterization presented here laser-treated and untreated alloy, thus providing the
confirms the consistency of the interfacial (corrosion mechanistic basis for the insignificant improvement in
product film/electrolyte and metal/electrolyte) electro- corrosion resistance due to laser surface melting.

METALLURGICAL AND MATERIALS TRANSACTIONS A VOLUME 44A, MAY 2013—2355


15 IV. CONCLUSIONS
1. Laser parameters have profound influence on the
12 resultant surface microstructure. The cooling rate is
pH of 0.001 M NaCl

higher at faster scan rates, and hence the degree of


9 microstructural refinement achieved in the case of
the faster scan rate is higher. A higher laser dwell
time leads to a relatively more homogeneous micro-
6
structure, which can be attributed to an increased
time in the liquid phase for dissolution of second
3 phases. Higher line energy and laser dwell time lead
to a higher melt depth.
0 2. A high laser power density and a faster scan rate
0 2 4 6 8 provide a high input energy and a rapid cooling
Time (day) rate, causing high thermal stress and cracking in the
laser-treated zone.
Fig. 15—Time-dependent change in the pH of 0.001 M NaCl in 3. Corrosion resistance of the slower scan rate speci-
which ZE41 samples were immersed for different days. men was similar to the untreated alloy, which can

(a)
Completely
corroded
grains

Pits

100 µm

(b)

Pits

100 µm

(c)

Pits

100 µm

Fig. 16—Corrosion morphology of the (a) untreated, (b) faster scan rate (sample 1), and (c) slower scan rate (sample 2) specimens at 55 h of
immersion in 0.001 M NaCl.

2356—VOLUME 44A, MAY 2013 METALLURGICAL AND MATERIALS TRANSACTIONS A


be attributed to their similar corrosion kinetics at 18. J.D. Majumdar, R. Galun, B.L. Mordike, and I. Manna: Mater.
the different interfaces (corrosion product film/solu- Sci. Eng. A, 2003, vol. 361, p. 119.
19. A.E. Coy, F. Viejo, F.J. Garcia–Garcia, Z. Liu, P. Skeldon, and
tion interface and the metal/corrosion product film G.E. Thompson: Corros. Sci., 2010, vol. 52 (2), p. 387.
interface). The laser-treated specimen produced with 20. Y.C. Guan, W. Zhou, and H.Y. Zheng: J. Appl. Electrochem.,
a faster scan rate had the lowest corrosion resis- 2009, vol. 39 (9), p. 1457.
tance, which is attributed to the presence of cracks 21. Y. Gao, C. Wang, M. Yao, and H. Liu: Mater. Corros., 2007,
vol. 58 (6), p. 463.
in the laser-treated zone of this specimen. 22. L.F. Guo, T.M. Yue, and H.C. Man: J. Mater. Sci., 2005, vol. 40
(13), p. 3531.
23. S.Y. Liu, J.D. Hu, Y. Yang, Z.X. Guo, and H.Y. Wang: Appl.
Surf. Sci., 2005, vol. 252, p. 1723.
24. G. Abbas, Z. Liu, and P. Skeldon: Appl. Surf. Sci., 2005, vol. 247
(1–4), p. 347.
ACKNOWLEDGMENTS 25. R.K.S. Raman, S. Murray, and M. Brandt: Surf. Eng., 2007,
vol. 23, p. 107.
Author, P. Chakraborty Banerjee, is grateful to 26. A. Koutsomichalis, L. Saettas, and H. Badekas: J. Mater. Sci.,
Monash University and Swinburne University of Technol- 1994, vol. 29, p. 6543.
ogy for providing the necessary infrastructure and CAST 27. D. Dubé, M. Fiset, A. Couture, and I. Nakatsugawa: Mater. Sci.
CRC for their funding support to carry out the project. Eng. A, 2001, vol. 299 (1–2), p. 38.
CAST was established under and is supported in part 28. J.D. Majumdar, T. Maiwald, R. Galun, B.L. Mordike, and I.
Manna: Laser. Eng., 2002, vol. 12 (3), p. 147.
by the Australian Government’s Cooperative Research 29. P.C. Banerjee, R.K.S. Raman, Y. Durandet, and G. McAdam:
Centre (CRC) Program. The authors would also like to Corros. Sci., 2011, vol. 53 (4), p. 1505.
acknowledge the Monash Centre for Electron Microscopy 30. Y. Durandet, S. Sun, and M. Brandt: Mater. Sci. Forum, 2010,
(MCEM) for providing the electron microscopy facility. vols. 654–656, p. 759.
31. S.K. Das, B.H. Kear, and C.M. Adams, eds.: Rapidly Solidified
Crystalline Alloys. 1985, The Metallurgical Society, Warrendale,
PA.
REFERENCES 32. A.S. Khanna, R.K. Singh Raman, E.W. Kreutz, and A.L.E.
Terrance: Corros. Sci., 1992, vol. 33 (6), p. 949.
1. N. Pebere, C. Riera, and F. Dabosi: Electrochim. Acta, 1990, 33. A. Conde, I. Garcia, and J.J. de Damborenea: Corros. Sci., 2001,
vol. 35, p. 555. vol. 43 (5), p. 817.
2. A. Pardo, M.C. Merino, A.E. Coy, R. Arrabal, F. Viejo, and E. 34. Y.S. Lim, H.P. Kim, J.H. Han, J.S. Kim, and H.S. Kwon: Corros.
Matykina: Corros. Sci., 2008, vol. 50 (3), p. 823. Sci., 2001, vol. 43 (7), p. 1321.
3. S.K. Das and L.A. Davis: Mater. Sci. Eng., 1988, vol. 98, p. 1. 35. M.C. Garcı́a-Alonso, M.L. Escudero, V. López, and A. Macı́as:
4. G. Song, A. Atrens, W. Xianliang, and B. Zhang: Corros. Sci., Corros. Sci., 1996, vol. 38 (3), p. 515.
1998, vol. 40 (10), p. 1769. 36. S. Virtanen, H. Böhni, R. Busin, T. Marchione, M. Pierantoni,
5. G. Song, A. Atrens, and M. Dargusch: Corros. Sci., 1998, vol. 41 and E. Blank: Corros. Sci., 1994, vol. 36 (9), p. 1625.
(2), p. 249. 37. K. Hashimoto, N. Kumagai, H. Yoshioka, J.H. Kim, E. Akiyama,
6. R.K.S. Raman: Metall. Mater. Trans. A, 2004, vol. 35A, p. 2525. H. Habazaki, S. Mrowec, A. Kawashima, and K. Asami: Corros.
7. M.C. Zhao, M. Liu, G.L. Song, and A. Atrens: Adv. Eng. Mater., Sci., 1993, vol. 35 (1–4 pt 1), p. 363.
2008, vol. 10 (1–2), p. 104. 38. American Society for Testing and Materials, ASTM G1-90, ASTM
8. W.C. Neil, M. Forsyth, P.C. Howlett, C.R. Hutchinson, and G1-90: Standard Practice for Preparing, Cleaning, and Evaluating
B.R.W. Hinton: Corros. Sci., 2009, vol. 51 (2), p. 387. Corrosion Test Specimens, 1999.
9. W.C. Neil, M. Forsyth, P.C. Howlett, C.R. Hutchinson, and 39. D.A. Jones, Principles and Prevention of Corrosion, 1996, Prentice
B.R.W. Hinton: Corros. Sci., 2011, vol. 53 (10), p. 3299. Hall, NJ, USA.
10. Y. Song, D. Shan, R. Chen, F. Zhang, and E.-H. Han: Corros. 40. X. Cao, M. Jahazi, J.P. Immarigeon, and W. Wallace: J. Mater.
Sci., 2009, vol. 51 (1), p. 62. Process Tech., 2006, vol. 171 (2), p. 188.
11. S. Izumi, M. Yamasaki, and Y. Kawamura: Corros. Sci., 2009, 41. F. Zucchi, V. Grassi, A. Frignani, C. Monticelli, and G.
vol. 51 (2), p. 395. Trabanelli: J. Appl. Electrochem., 2006, vol. 36, p. 195.
12. G.L. Makar and J. Kruger: J. Electrochem. Soc., 1990, vol. 137 (2), 42. G. Baril and N. Pébère: Corros. Sci., 2001, vol. 43 (3), p. 471.
p. 414. 43. E. Barsoukov and J.R. Macdonald, eds.: Impedance Spectroscopy:
13. M. Qian, D. Li, and C. Jin: Sci. Technol. Adv. Mat., 2008, vol. 9 Theory, Experiment, and Applications, 2nd ed. Wiley, Hoboken,
(2), p. 025002. 2005.
14. P. Volovitch, J.E. Masse, A. Fabre, L. Barrallier, and W. Saikaly: 44. M.E. Orazem and B. Tribollet: in Electrochemical Impedance
Surf. Coat. Tech., 2008, vol. 202 (20), p. 4901. Spectroscopy, Wiley, Hoboken, 2008.
15. Y. Gao, C. Wang, Q. Lin, H. Liu, and M. Yao: Surf. Coat. Tech., 45. W. Liu, F. Cao, A. Chen, L. Chang, J. Zhang, and C. Cao: Corros.
2006, vol. 201 (6), p. 2701. Sci., 2010, vol. 52 (2), p. 627.
16. R. Subramanian, S. Sircar, and J. Mazumdar: J. Mater. Sci., 1991, 46. F. Zucchi, A. Frignani, V. Grassi, A. Balbo, and G. Trabanelli:
vol. 26, p. 951. Mater. Chem. Phys., 2008, vol. 110 (2–3), p. 263.
17. A.A. Wang, S. Sircar, and J. Mazumder: J. Mater. Sci., 1993, 47. P. Chakraborty Banerjee and R.K. Singh Raman: Electrochim.
vol. 28, p. 5113. Acta, 2011, vol. 56 (11), p. 3790.

METALLURGICAL AND MATERIALS TRANSACTIONS A VOLUME 44A, MAY 2013—2357

S-ar putea să vă placă și