Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Materials and Design 57 (2014) 40–50

Contents lists available at ScienceDirect

Materials and Design


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

Corrosion behaviour of laser surface melted magnesium alloy AZ91D


C. Taltavull a, B. Torres a, A.J. Lopez a, P. Rodrigo a, E. Otero a, A. Atrens b,⇑, J. Rams a
a
Dpto. Ciencia e Ingeniería de Materiales, ESCET, Universidad Rey Juan Carlos, C/Tulipán s/n, Móstoles Madrid 28933, Spain
b
The University of Queensland, Division of Materials, St. Lucia, Qld 4072, Australia

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

Article history: A high power diode laser (HPDL) was used to produce laser surface melting (LSM) treatments on the
Received 22 October 2013 surface of the Mg alloy AZ91D. Different treatments with different microstructures were produced by
Accepted 28 December 2013 varying the laser-beam power and laser-scanning speed. Corrosion evaluation, using hydrogen evolution
Available online 3 January 2014
and electrochemical measurements, led to a relationship between microstructure and corrosion. Most
corrosion rates for LSM treated specimens were within the scatter of the as-received AZ91D, whereas
Keywords: some treatments gave higher corrosion rates and some of the samples had corrosion rates lower than
Magnesium
the average of the corrosion rate for AZ91D. There were differences in corroded surface morphology.
Laser processing
Polarization
Nevertheless laser treatments introduced surface discontinuities, which masked the effect of the micro-
Scanning electron microscopy structure. Removing these surface defects decreased the corrosion rate for the laser-treated samples.
Ó 2014 Elsevier Ltd. All rights reserved.

1. Introduction aluminium alloys using LSM treatments. Padovani et al. [37]


eliminated the pitting corrosion which occurred after friction stir
Magnesium (Mg) alloys are of interest to the automotive and welding, using LSM as a post-treatment for repairing sensitisation
aerospace industries because of their low density [1] and high in the weld region. Li et al. [38] increased the corrosion fatigue life
specific strength. Moreover, Mg alloys are gaining interest for biode- of an aluminium alloy by a factor of between two and four times.
gradable medical implants as a consequence of a good combination Viejo et al. [39] obtained an increase in intergranular corrosion re-
of load bearing mechanical strength, an elastic modulus close to that sistance of LSM treated aluminium alloys, although the presence of
of bone, biocompatibility and an ability to promote osteogenesis defects and solute segregation bands promoted galvanic corrosion
[2–4]. Nevertheless, their poor resistance to corrosion, and wear, at the interface between the melted layer and the bulk alloy. Rams
limits wider application [1,5–8]. Mg corrosion behaviour is signifi- et al. [40,41] improved the corrosion resistance of aluminium
cantly influenced by the microstructure [9–21]. Second phases alloys and aluminium matrix composites reinforced with SiC parti-
cause micro-galvanic couples increasing the corrosion rate. Laser cles [40,41]. The corrosion was improved, although the presence of
surface melting (LSM) can homogenise and refine the microstruc- ceramic particles limited the efficiency of the improvement.
ture, and can dissolve the second phases. These microstructure These successful LSM treatments indicated that LSM treatments
changes could improve the corrosion behaviour [22–26] and wear could also improve the corrosion behaviour of Mg alloys. Mondal et
behaviour [27–29] of Mg alloys. al. [22] achieved an increase in the corrosion resistance by LSM
LSM treatments have been investigated for the most common treatments of the Mg alloy ACM720. The polarization resistance
structural metals: steels and aluminium alloys. Kwok et al. of the laser-treated sample was double that of the as-cast alloy.
[30–32] developed LSM for the desensitization of austenitic and Khalfaoui et al. [23] developed a LSM treatment for ZE41, after
duplex stainless. They achieved a complete removal of the chro- which the LSM treated sample was unaltered after 130 hours in
mium carbides, and thereby improved the intergranular corrosion the salt-spray corrosion test, whereas the as-received ZE41 was
resistance. Bao et al. [33] used LSM treatment to repair stress completely corroded. Coy et al. [24] obtained passive behaviour
corrosion cracking in Inconel alloy 182, by successfully sealing ex- for AZ91D after LSM treatment. Abbas et al. [25] decreased the
isting stress corrosion cracks. Li et al. [34], Chong et al. [35] and corrosion rate of WE43 by up to 87%. In all this research on Mg
Osorio et al. [36] improved the resistance to pitting corrosion of alloys [22–25], the improvement of the corrosion behaviour was
attributed to the refinement in the microstructure, and the
increased solubility of the a-Mg matrix.
⇑ Corresponding author. Tel.: +61 7 3365 3748. In contrast, Chakraborty et al. [26], in an Electrochemical
E-mail address: andrejs.atrens@uq.edu.au (A. Atrens). Impedance Spectroscopy (EIS) study of LSM treated ZE41, did not

http://dx.doi.org/10.1016/j.matdes.2013.12.069
0261-3069/Ó 2014 Elsevier Ltd. All rights reserved.
C. Taltavull et al. / Materials and Design 57 (2014) 40–50 41

observe any significantly change in the corrosion behaviour,


although microstructural refinement was achieved. Liu et al. [42]
reported other exceptions in their study of LSM treatments on
aluminium alloys.
The LSM treatments have used laser sources such as Nd:YAG
[22,25,30–33,36], excimer [23,24,35,37,39] or CO2 [35,38]. How-
ever, the high power diode laser (HPDL) has advantages, such as
better compactness, energy efficiency, lifetime and running costs.
In addition, new selective laser surface melting (SLSM) treatments
can be developed using this laser source [43]. Consequently, its use
in the field of laser materials processing has been increasingly
recognized [40,41,44].
The overarching aim of this research was to study the influence
of processing parameters on the corrosion behaviour of LSM AZ91D
using a HPDL. AZ91D is of particular interest as the most common
industrial Mg alloy. This aim was addressed by seeking to under-
stand the relationship between (i) the microstructure, (ii) the
Fig. 1. Schematic section of the electrochemical cell used for the electrochemical
LSM parameters laser beam power and scanning speed, and (iii)
tests. The sample was the working electrode (WE). RE is the reference electrode, and
corrosion rates measured by hydrogen evolution and polarization CE is the counter electrode.
curves. Moreover, this research also evaluated the influence of
other important parameters such as surface roughness on the
The laser melted surface before and after the corrosion evalua-
corrosion behaviour of the LSM treated specimens.
tion was characterized using a Scanning Electron Microscope (SEM,
Hitachi S-3400N) equipped with an Energy Dispersive X-ray Spec-
2. Experimental procedure
trometer (EDS, XFlash 5010 Bruker). The accelerating voltage was
10 or 15 kV. The working distance was about 10 mm.
2.1. Materials
The influence of the surface roughness, surface defects and
graphite absorption of the LSM treated specimens was evaluated
AZ91D was supplied by Magnesium Elektron Ltd and had the
by lightly grinding the surface of some of the LSM treated speci-
following composition (in weight %): 9% Al, 0.67% Zn, 0.23% Mn
mens to 1200 grit, and carrying out electrochemical evaluations
and balance Mg. The material was supplied as billets of 250 mm
of these LSM treated and ground samples, and comparing these re-
diameter, and was cut into specimens of 30 mm  25 mm  3
sults with those of the as-treated samples. This procedure allowed
mm for the laser treatments. Due to the fact that the metallic
evaluation of the influence of: (i) surface roughness of the LSM
surface is highly reflective to the laser radiation, each specimen
treated surface, (and the associated higher real surface area
surface was painted with Aerodag Ò G (supplied by EMS, USA) to
exposed to the corrosive medium of the LSM treated surface), (ii)
improve the absorption of the laser radiation into the specimen,
carbon incorporation into the surface of the LSM treated surface,
and to avoid its oxidation during the laser treatment. AerodagÒ G
and (iii) superficial voids. Care was taken to ensure that the
(EMS, USA) is a commercial dry-film lubricant of specialty
grinding removed the surface roughness induced by the laster
processed micro-sized graphite in isopropyl alcohol. The paint
processing, and that the laser surface melted microstructure was
was applied by spraying one coat over the specimen, which quickly
retained for study.
dried, obtaining a black homogeneous carbon coating of about 10.2
± 3.5 lm of thickness.
2.4. Electrochemical measurements
2.2. Laser treatment
Electrochemical measurements were carried out in 3.5 wt.%
The LSM treatments were carried out using a Rofin-Dilas DS13 NaCl at 24 ± 3 ºC without stirring, using a specimen with an area
HPDL, controlled with an ABB IRB 2400/16 robot. The laser spot of 0.76 cm2 exposed to the solution in a conventional three-
was scanned over the surface of the specimen. The laser electrode cell as shown in Fig. 1, with the horizontal specimen as
beam-power was varied in the range of 375–600 W, and the laser the working electrode (WE), a silver/silver chloride (Ag/AgCl) refer-
scanning-speed was 45–60–90 mm s1. The distance between ence electrode (RE) saturated with KCl, a graphite rod as counter
consecutive scans was 1.55 mm in all cases, and the distance electrode (CE), and using an Autolab PGStat 30 potentiostat.
between the laser and the specimen surface was 7 mm. Polarization curves were measured by scanning from –400 to
The laser input energy (EL) was calculated as follows: 800 mV with respect to the open circuit potential (OCP) using a
1 mV s1 scan rate. No iR compensation was applied to the anodic
P
EL ¼ ð1Þ polarisation curves (where i is the current and R is the resistance
v between the reference electrode and the specimen). The polarisa-
where P is the laser beam power, and v the laser scanning speed. tion curves were reported as measured. The amount of iR compen-
sation was negligible in the parts of the polarisation curves that
2.3. Specimen characterization were used to evaluate the Tafel constants. The Tafel extrapolation
of the cathodic branch of the polarization curve between OCP 30
After LSM treatments and corrosion tests, each specimen was mV and OCP 80 mV was used to evaluate the corrosion current
analysed in cross-section by mounting in an electrically conductive density, icorrTafel, using the software of GPES 4.8 (by Eco Chemie
resin, wet grinding using a sequence of silicon carbide abrasive B.V, The Netherlands). The cathodic polarisation curve was
papers from 120 to 2400 grit, and polishing with 0.05 lm diamond approximately linear in this potential range (on a linear-log plot
paste using ethylene glycol as lubricant. The samples were etched as usually used, see e.g. Fig. 8). Kirkland et al. [45] have recently
using a solution of 20 ml acetic acid, 1 ml nitric acid, 60 ml shown that such Tafel extrapolation is the most widely used
ethylene glycol and 20 ml distilled water. method of measuring the corrosion rate for Mg alloys.
42 C. Taltavull et al. / Materials and Design 57 (2014) 40–50

Fig. 3. (a) Macroscopic view of LSM treated AZ91 at laser beam power 375 W and
laser scanning speed 90 mm s1. (b) Representative SEM micrograph of a cross-
section, showing that the laser treatment had modified only the b-phase. (c) Higher
magnification of (b). The microstructure shown corresponds to the laser treatment
named selective laser surface melting (SLSM).

Fig. 2. SEM micrograph of the as-received Mg alloy AZ91 showing second phase
(b-Mg17Al12) along dendrite boundaries.

Although the corrosion morphology was somewhat non-


uniform, it is nevertheless useful to present the corrosion rate data
as a numeric value with the same units for the different
measurement techniques, so that the results for each measure-
ment technique can be directly compared. The instantaneous
corrosion rate, Pi (mm y1) was determined from the corrosion
current density, icorr using [46,47]:
Pi ¼ 22:85icorr ð2Þ Fig. 4. (a) Macroscopic view of LSM treated AZ91 at laser beam power 488W and
laser scanning speed 60 mm s1. (b) Representative SEM micrograph of a cross-
section showing the microstructure for this intermediate input energy. (c) Higher
2.5. Hydrogen evolution magnification of (b).

Immersion test were carried out in 3.5 wt.% NaCl, exposing a


specimen area of approximately 3 cm2, at 24 ± 3 ºC and without
stirring. The evolved hydrogen was collected into a burette by a
funnel above the corroding specimen as described by Song et al.
[48]. For the immersion test of the LSM specimens, an isolating
coating was applied to the non-treated surface to ensure that there
was corrosion only at the LSM treated area. Hydrogen evolution al-
lows the measurement of the average corrosion rate over the
whole immersion period, and also the instantaneous corrosion rate
during the immersion test. The hydrogen evolution rate VH (mL
cm2 d1) was used to evaluate the corrosion rate, PH (mm y1)
by [46,47]:
Fig. 5. (a) Macroscopic view of LSM treated AZ91 at laser beam power 600 W and
PH ¼ 2:088V H ð3Þ
laser scanning speed 45 mm s1. (b) Representative SEM micrograph of a cross-
section. (c) Higher magnification of (b).

3. Results
matrix. Higher magnification of the laser-modified b-phase pre-
3.1. Microstructure and surface characterization sented in Fig. 3c shows an eutectic microstructure based on fine
plates of b-Mg17Al12 (light grey) and a-Mg phase (dark grey). The
Fig. 2 shows that the microstructure of the as-received AZ91D microstructure shown corresponds to the laser treatment named
consisted mainly of the Mg matrix (a-Mg) (dark grey areas) and selective laser surface melting (SLSM) [43,49]. Fig. 4 presents the
the second b-phase (Mg17Al12) (lighter grey areas). EDS mapping results using the laser beam power 488 W and a laser scanning
showed the distribution of the main alloying elements (not speed of 60 mm s1. The surface was homogenous with no rem-
shown). nants of graphite paint and there were some localized voids (Fig.
Figs. 3–5 present the typical appearance of the surface after sur- 4a). The cross-section SEM micrograph (Fig. 4b) indicated that
face modifications by LSM, and the corresponding microstructure the microstructure for this intermediate input energy (Table 1) in
observed by SEM cross-sections at different magnifications. Fig. 3 the modified surface layer consisted of fine plates of b-Mg17Al12
corresponds to LSM using 375 W laser beam power and 90 mm s1 phase (light grey) and an a-Mg matrix (dark grey) distributed
laser scanning speed, the lowest input energy used (Table 1). This randomly along the surface, producing an homogeneous and
LSM produced almost no deformation of the treated surface, continuous modified layer (Fig. 4c). In addition, there was some
although the location of the laser treated surface was clear un-melted a-Mg matrix, indicated by the surrounding lines in
(Fig. 3a). There was no remnant of the graphite paint on the laser Fig. 4b. The treated depth was approximately 300 lm. Fig. 5a
treated surface. The SEM micrograph of the cross-section, shown shows that, for the highest input energy (Table 1), i.e. laser beam
in Fig. 3b, indicated that the low input energy caused only modifi- power 600 W and laser scanning speed of 45 mm s1, there was
cation of the b-Mg17Al12 phase without any change in the a-Mg a large deformation of the surface. The complete removal of the
C. Taltavull et al. / Materials and Design 57 (2014) 40–50 43

Table 1
Data of laser input energies for the different laser treatments.

Specimen Laser beam power (W) Laser scanning speed (mm s1) Laser input energy (J m1)
AZ91D
L-AZ91D-375–90 375 90 4.17  103
L-AZ91D-488–90 488 90 5.42  103
L-AZ91D-600–90 600 90 6.67  103
L-AZ91D-375–60 375 60 6.25  103
L-AZ91D-488–60 488 60 8.13  103
L-AZ91D-600–60 600 60 1.00  102
L-AZ91D-375–45 375 45 8.33  103
L-AZ91D-488–45 488 45 1.08  102
L-AZ91D-600–45 600 45 1.33  102

graphite paint did not occur. The cross-section SEM micrograph


(Fig. 5b) showed a microstructure composed of a fine a-Mg matrix
and b-Mg17Al12. There was no un-melted a-Mg matrix. Higher
magnification, presented in Fig. 5c, revealed that the microstruc-
ture was characterized by very small isolated b-Mg17Al12 phase
particles immersed in a continuous aluminium rich a-Mg matrix.
The treated depth was approximately 600 lm.

3.2. Immersion tests

Fig. 6 presents the specific hydrogen evolution volume for the as-
received AZ91D, and the LSM treated samples, as a function of the
immersion time. For all specimens, the hydrogen volume increased
with immersion time. For the as-received AZ91D, the error bars in-
dicate the data variability, as deduced from the three samples tested.
Fig. 6b (magnification of the squared zone in Fig. 6a) indicates that,
with the exception of the lowest laser power, the hydrogen evolu-
tion rate for short immersion times for all the LSM samples was lar-
ger than that for the as-received AZ91D. Fig. 6a indicates that none of
the LSM samples produced a hydrogen evolution rate lower than
that of the as-received AZ91D, taking into consideration the error
bars for the AZ91D. Furthermore, the highest laser-beam powers
produced LSM had hydrogen evolution rates significantly higher
than of the as-received AZ91D. However, as the immersion time
increased, the differences between the hydrogen volume for the
as-received AZ91D and the LSM treated samples reduced, and only
the LSM samples treated at the highest input energies corroded
faster that the as-received AZ91D. The degradation of the other
LSM samples was within the variability of the degradation of the
as-received AZ91D. Some of these samples had corrosion rates lower
than the average corrosion rate for AZ91D.
Fig. 7 presents contour plots of the average corrosion rate
related to laser beam power and laser scanning speed. The black
colour zone represents the average corrosion rate for the as-
received AZ91D, which was 1.3 ± 0.8 mm y1 averaged over the
Fig. 6. Hydrogen evolution data for AZ91D and LSM specimens immersed for 50
first 24 hours and 2.6 ± 1.4 mm y1 averaged over the first 48
hours in 3.5% NaCl. For the LSM samples, the designation gives the laser power and
hours. scanning speed. (b) Detail view of dashed square in (a). The data for AZ91D are
The corrosion rate after 24 hours (Fig. 7a) indicates a correlation given as the full squares, and the error bars indicate data variability. Most of the
between laser parameters and corrosion rate of the LSM treated LSM samples had corrosion rates higher than or within the scatter of the AZ91D
samples. The corrosion rate increased with increasing laser-beam data. At the longer immersion times, some of the samples had corrosion rates lower
than the average of the corrosion rate for AZ91D.
power and decreasing laser-scanning speed. The lowest corrosion
rates corresponded to (i) the lowest laser-beam power (375 W)
and highest laser-scanning speed (90 mm s1), which correspond speeds. After 48 hour most of the lowest input energies laser treat-
to the lowest input energy and SLSM and (ii) middle laser-beam ments provided corrosion rates similar to those of as-received
power (450 W) and middle laser-scanning speed (60 mm s1). AZ91D, and some conditions provided corrosion rates lower than
The highest corrosion rate corresponded to the highest input those of as-received AZ91D.
energies, i.e. 600 W laser-beam power and 45 mm s1 laser-
scanning speed. 3.3. Polarization curves
Fig. 7b indicates that there was a similar correlation for the cor-
rosion rate averaged over 48 hour. The corrosion rate increased Fig. 8 presents typical polarization curves measured after 1 hour
with high laser-beam power and low laser-scanning speeds, and immersion in 3.5% NaCl solution for the as-received AZ91D and
was lowest for low laser-beam power and high laser-scanning LSM samples. Fig. 8a presents the polarization curves for a
44 C. Taltavull et al. / Materials and Design 57 (2014) 40–50

Fig. 7. Contour maps of average corrosion rate evaluated from hydrogen evolution
data averaged (a) over 24 hours, and (b) 48 hours of immersion in 3.5% NaCl related
to laser-beam power and laser-scanning speed. Note that, at the longer immersion
times, some of the samples had corrosion rates lower than the average of the
corrosion rate for AZ91D.

laser-scanning speed of 90 mm s1 and laser-beam powers of 375–


488–600 W. For the combination of laser beam power of 375 W
and laser scanning speed of 90 mm s1 (SLSM treatment) the polar-
ization curve moved slightly to the left, which meant a slight
decrease in the corrosion current density, icorr. For the other
laser-beam powers, as the laser-beam power increased, the curves
moved to higher values of OCP, but they also moved to the right so
the corrosion current density, icorr, increased with increasing laser-
beam power. Fig. 8b presents the polarisation curves for the laser-
scanning speed of 60 mm s1. All curves for laser treated samples
had values of icorr higher than that of the as-received AZ91D and Fig. 8. Polarization curves after one hour immersion in 3.5% NaCl for AZ91D-LSM
icorr increased with increasing laser beam power. There was the samples: (a) related to laser-beam power for a laser scanning speed of 90 mm s1
same tendency for a laser scanning-speed of 45 mm s1 (not (b) related to laser-beam power for a laser scanning speed of 60 mm s1, and (c)
shown). related to laser-scanning speed for a laser-beam power of 375 W.
Fig. 8c presents the polarization curves for a laser-beam power
of 375 W and laser scanning speeds of 90–60–45 mm s1. As the Fig. 9 presents a contour plot of the corrosion rate calculated
laser scanning speed decreased, a displacement to the right of from the Tafel extrapolation method related to laser-beam power
the polarization curves occurred. The equivalent tendency and laser-scanning speed. The corrosion rate for as-received
occurred when laser beam power were fixed at 488W-600 W AZ91D alloy was 0.46 ± 0.16 mm y1, which is represented on
(not shown). the plots by the black colour. The corrosion rate increased with
C. Taltavull et al. / Materials and Design 57 (2014) 40–50 45

Table 2
Data of corrosion rates for AZ91D-LSM samples in 3.5% NaCl measured by immersion test and electrochemical test represented in Figs. 6 and 8.

Specimens Laser beam Laser scanning Ecorr (VAg/AgCl) icorrTafel (lA cm2) PiTafel (mm y1) PAH (24 h) (mm y1) PAH (48 h) (mm y1)
power (W) speed (mm s1)
AZ91D 1.518 20.2 ± 7.2 0.46 ± 0.16 1.3 ± 0.8 1.6 ± 1.4
L-AZ91D-375–90 375 90 1.527 12.2 0.28 0.9 1.3
L-AZ91D-488–90 488 90 1.522 108.9 2.49 4.0 4.3
L-AZ91D-600–90 600 90 1.480 379.9 8.68 4.1 2.0
L-AZ91D-375–60 375 60 1.524 54.6 1.25 1.8 0.9
L-AZ91D-488–60 488 60 1.488 161.4 3.69 4.9 2.4
L-AZ91D-600–60 600 60 1.488 621.5 14.20 6.1 5.2
L-AZ91D-375–45 375 45 1.523 95.2 2.18 3.1 1.8
L-AZ91D-488–45 488 45 1.481 260.9 5.96 6.6 6.2
L-AZ91D-600–45 600 45 1.500 569.1 13.00 20.9 17.7

Fig. 11. Macroscopic image of the corroded surface of as-treated AZ91D and
AZ91D-LSM samples after measurement of the polarization curve in 3.5% NaCl,
related to laser-beam power and laser-scanning speed. The cell configuration (Fig.
1) produced circular corroded zones, which were the areas in focus.

Fig. 9. Contour map of corrosion rate calculated using Tafel extrapolation method
from polarization curves at one hour immersion in 3.5% NaCl related to laser-beam presented a slightly lower corrosion rate. The data of Fig. 9 are in
power and laser-scanning speed. broad agreement with those of Fig. 7.
Fig. 10 and Table 2 present a comparison between the corrosion
rate evaluated by the Tafel extrapolation and from the evolved
hydrogen averaged over 24 and 48 hours of immersion. These
quantities should be equal so the dotted line is an aid to the eye
and shows the condition for equality. There was an agreement be-
tween the corrosion rates obtained by the Tafel extrapolation and
the hydrogen evolution methods, although there was significant
data scatter.

3.4. Corroded surface characterization

Fig. 11 presents macroscopic views of the corroded surface of


the as-received AZ91D, and typical LSM treated samples after the
polarization tests in relation to laser-beam power (P) and laser-
scanning speed (v). The cell configuration (Fig. 1) produced circular
corroded zones, which were the areas in focus. There was an in-
crease of the surface roughness produced by the LSM treatments
as the laser input energy increased, i.e. at high laser-beam power
and low laser-scanning speed.
Fig. 10. Comparison between the corrosion rates for immersion in 3.5% NaCl The AZ91D corroded surface was characterized by large pits
measured by hydrogen evolution averaged at 24 and 48 hours of immersion and distributed throughout the surface. For LSM treated samples,
Tafel extrapolation method. the corroded surface morphology depended on the laser para-
meters. For the lowest input energies, SLSM treatments, the cor-
roded surface was similar to that of the as-received AZ91D. At
increasing laser-beam power and decreasing laser-scanning speed. medium input energies, i.e. laser beam power of 488 W, the
Moreover, all LSM treatments presented a corrosion rate higher corroded surface of the LSM treated sample depended on the
than the as-received AZ91D, with the exception of the condition laser-scanning speed. At the highest laser scanning speed (90
corresponding to the lowest input energy, SLSM, (laser beam mm s1) there was pitting on the corroded surface, although
power of 375 W and laser scanning speed of 90 mm s1), which the pit distribution did not seem to be random, but the pits were
46 C. Taltavull et al. / Materials and Design 57 (2014) 40–50

Fig. 12. (a) Cross-section backscattered SEM micrograph of the corroded as


received AZ91D after 54 hour immersion in 3.5% NaCl, (b) magnification of (a).

Fig. 13. (a) Cross-section backscattered SEM micrograph of the corroded surface of
AZ91D-SLSM sample after 54 hour immersion in 3.5% NaCl; (b) arrowed detail 1 in
(a); (c) arrowed detail 2 in (a).

oriented preferentially in the direction of laser scanning. At the


laser-scanning speed of 60 mm s1, the corroded surface pre-
sented a different morphology. There were no pits, but there Fig. 14. (a) Cross-section backscattered SEM micrograph of the corroded surface of
was significant degradation oriented in the direction of move- AZ91D-LSM sample treated at laser-beam power 450 W and laser-scanning speed
ment of the laser spot. At the lowest laser-scanning speed, i.e. 60 mm s1 after 54 hour immersion in 3.5% NaCl; (b) magnified view of dashed
white square in (a); (c) EDS line scan indicated in (b); (d) EDS-line scan presented in
45 mm s1, there was high degradation of the surface with a large
(c) without Mg data.
number of voids, corresponding to an increased amount of
corrosion. There was similar corrosion of the corroded surface
of LSM treated samples at the highest input energies, i.e. laser
beam power of 600 W combined with a large number of voids.

3.5. Cross-sections of corroded surfaces

Figs. 12–15 present typical cross-sections of the corroded


surfaces of as-received AZ91D, and LSM treated samples. Fig. 12
presents a cross-section SEM micrograph of as-received AZ91D
after 54 hours of immersion. Fig. 12a indicates that the corrosion
progressed as dissolution of the a-Mg phase (dark grey) whilst
Fig. 15. (a) Cross-section backscattered SEM micrograph of the corroded surface of
the b-Mg17Al12 phase (light grey) was unaltered (Fig. 12b). More- AZ91D-LSM sample treated at laser-beam power 600 W and laser-scanning speed
over, Fig. 12a shows, in agreement with the literature [35], that 60 mm s1 after 1 hour immersion in 3.5% NaCl; (b) Detail of (a); (c) Dashed square
b-Mg17Al12, in a continuous network, could act as a barrier to detail in (b) and (d) full square detail in (b).
prevent the progress of corrosion. A detail of this phenomenon is
presented in Fig. 12b.
Fig. 13 presents cross-section SEM micrographs, after 54 hours Fig. 14 presents a cross-section SEM micrograph of the LSM
of immersion, of the SLSM samples, i.e. laser-power of 375 W and treated sample with medium input energy, i.e. laser-beam power
laser-scanning speed of 90 mm s1. The morphology of the corro- of 450 W and laser-scanning speed of 60 mm s1. Fig. 14a indicated
sion was similar to that of the as-received AZ91D. Fig. 13a shows the presence of homogeneous corrosion and the absence of pitting
that there were two different corrosion morphologies in the corrosion. Fig. 14b (a magnification of the dashed square in Fig.
cross-section corroded surface of the SLSM treated sample: (i) pits, 14a) indicated the formation of a thin film on the surface of the
indicated in Fig. 13a by arrow 1, and (ii) dissolution of the laser LSM sample. In addition, Fig. 14b revealed the existence of some
modified b-phase, indicated in Fig. 13a by arrow 2. Fig. 13b cracks in the thin surface layer (as indicated by the circles). EDS
presents a magnification of an area similar to that designated by line scan performed as indicated by the line in Fig. 14b (Fig. 14c)
the arrow 1 in Fig. 13a, where there was a deep pit next to the la- indicated that the surface layer was composed mainly of magne-
ser-modified b-phase. Fig. 13c (magnification of area identified by sium and oxygen. Fig. 14d presents a magnification of the line scan
arrow 2 in Fig. 13a) revealed another corrosion behaviour: homo- after removal of the Mg signal. Fig. 14d indicates that there was
genous dissolution of the laser-modified b-phase attributed to significant aluminium in the thin surface layer.
corrosion products and a slight depletion of the surface in the laser Fig. 15 presents a typical cross-section SEM micrograph, after 1
modified b-phase. hour of immersion, of high input energy LSM, i.e. laser-beam power
C. Taltavull et al. / Materials and Design 57 (2014) 40–50 47

10 showed that the corrosion rate data were consistent with the
expected linear relation between the corrosion rate evaluated from
the evolved hydrogen after 24 or 48 hours and the corrosion rate
evaluated by Tafel extrapolation of the polarization curves,
although there was significant scatter of the data. The contour
plots of Figs. 7 and 9 had many similarities, particularly for the
medium laser power. This analysis indicated that Tafel extrapola-
tion at short immersion times could be used as an indication of
the long-term corrosion behaviour, and that there were none of
the other issues that can cause Tafel extrapolation to give values
not in agreement with independent measurements [44,48,50].
The nature of the laser treatment applied had a strong influence
in the corrosion rate of the laser treated samples, in agreement
with [17]. In general, the corrosion rate increased with increasing
laser input energy, so that the corrosion rate was maximum for
high laser power and low laser-scanning speed (Fig. 7). An analysis
of the complete set of data for the two tested times, indicated that
the relationship between the laser parameters with the corrosion
Fig. 16. Polarization curves of as-received AZ91D and AZ91-LSM at laser power- rate was similar. However, at 48 hours of immersion time (Fig.
beam 600 W and laser-scanning speed 45 mm s1 measured as treated or ground to
1200 grit, in 3.5% NaCl.
7b) there was a significant area of low corrosion rates (represented
by the black area) at intermediate input energy which was not pre-
sent at 24 hours (Fig. 7a). This indicated that the corrosion rate
of 600 W and laser-scanning speed of 60 mm s1. Fig. 15a indicates tended to somewhat stabilize with longer immersion time. In addi-
pits distributed throughout the surface in the cross-section SEM tion, there were laser treatment conditions (black with white lines
micrograph of the LSM corroded sample. The pits were located pre- in Fig. 7b) where the corrosion rate was significantly lower than
ferentially in the defects (voids). The progression of the pitting the average of the as-received alloy, although the scatter of the
hardly depended on the microstructure achieved by the LSM treat- data was not included in this plot. Nevertheless, the fact that this
ment. Fig. 15b shows a large pit that had progressed through the region existed, indicates that this laser conditions needs to be
surface to a depth of 30 lm. There was an intermediate zone where taken into account.
the pitting had not progressed (indicated by the square in Fig. 15b). The differences on the corrosion behaviour of the laser-treated
Fig. 15c presents this zone at higher magnification. There was a specimens have been caused by two main factors: (i) the changes
continuous network of fine b-Mg17Al12, indicated by the arrows in the microstructure by means of the laser treatment [17,22–24]
in Fig. 15c. Nevertheless, Fig. 15d, (corresponding to a magnifica- and (ii) the modification produced in the surface characteristics,
tion of the dashed square in Fig. 15b) indicates that, in other including composition, roughness [51] or residual stress [52].
LSM zones, the microstructure was composed of isolated fine These are analysed separately in the following.
b-Mg17Al12 phase. Figs. 3c, 4c and 5c presented the typical SEM micrographs of the
three main microstructures obtained by the LSM treatments
[53]. Fig. 3c showed the microstructure modification achieved in
3.6. Influence of other parameters
a SLSM treatment (laser-beam power of 375 W and laser-scanning
speed of 90 mm s1) where only the b-Mg17Al12 phase was
Fig. 16 presents polarization curves of (i) as-received AZ91D, (ii)
modified and the a-Mg matrix was unaltered. The laser-modified
a LSM treatment as treated, and (iii) a LSM treatment ground to
b-phase presented a eutectic microstructure based on fine plates
1200 grit after the laser treatment. A displacement of the curve
of b-Mg17Al12 and a-Mg phase [43,49]. The corrosion behaviour
to the left occurred for the ground LSM sample corresponded to a
of the SLSM treatment was pitting combined with slight homoge-
lower corrosion current density, icorr.
neous dissolution of the laser modified b-phase (Fig. 13). This
Fig. 17 shows the corrosion rate calculated by Tafel extrapola-
corrosion behaviour could be explained as a consequence of: (i) a
tion from the polarisation curves for (i) the LSM samples tested
two-phase microstructure similar to that of the as-received
as-treated, and (ii) after grinding to 1200 grit. Table 3 presents
AZ91D, which promoted the formation of micro-galvanic couples
the numeric data for the corrosion rates measured related to the
between the a-Mg and the laser modified b-phase, causing
surface preparation. At the highest input energy, i.e. laser-beam
inhomogeneous corrosion, and (ii) the eutectic microstructure
power of 600 W and laser-scanning speed of 45 mm s1, there
produced by the laser modified b-phase was composed of fine
was a reduction of up to 97% of the corrosion rate after surface
b-Mg17Al12 phase forming a closed network, which implied a pro-
grinding, whereas there was a reduction of 45% for a middle input
tective barrier against the progression of the corrosion attack [35],
energy, i.e. laser-beam power of 488 W and laser-scanning speed of
so if some defect in the surface initiated corrosion, progression
90 mm s1. In addition, the corrosion rates of some laser-treated
along the laser modified b-phase zone was expected to be delayed.
specimens, i.e. laser beam power of 600 W and laser scanning
In the SLSM treatment, the corrosion rate was slightly lower than
speed of 60 mm s1, were lower than those of the as-received
that for the as-received AZ91D, probably due to the reduction of
AZ91D Mg alloy (Table 3).
the number of pits formed, as a consequence of combination of
pitting corrosion (Fig. 13b) and a homogenous dissolution of laser
4. Discussion modified b-phase (Fig. 13c), which is expected to avoid the
formation of a pit in the adjacent a-Mg matrix.
Fig. 6 showed that the hydrogen evolution data indicated that Fig. 4c presented a cross-section of the LSM treatment achieved
the corrosion rate of the as-received AZ91D and the LSM-treated using medium laser input energy, i.e. laser-beam power of 450 W
specimens was essentially constant with immersion time for short and laser-scanning speed of 60 mm s1, for which there was a low
immersion times. For longer times, corrosion progressed less in the corrosion rate (Fig. 7b). The microstructure consisted of fine plates
laser treated samples than in the as-received AZ91D (Fig. 6a). Fig. of b-Mg17Al12 and a a-Mg matrix distributed in a modified surface,
48 C. Taltavull et al. / Materials and Design 57 (2014) 40–50

Fig. 17. Comparison between the corrosion rates calculated for LSM specimens in as-treated and ground to 1200 grit surface conditions in 3.5% NaCl.

Table 3
Data of corrosion rates of AZ91D-LSM samples in 3.5% NaCl measured in as-treated and ground to 1200 grit surface condition represented in Fig. 17.

Specimen Laser beam power (W) Laser scanning speed Ecorr (VAg/AgCl) icorrTafel (lA cm2) PiTafel (mm y1)
(mm s1)
AZ91D 1.518 20.2 ± 7.2 0.46 ± 0.16
L-AZ91D-600–45– 600 45 1.504 41.4 0.90
1200G
L-AZ91D-375–60– 375 60 1.524 11.5 0.20
1200G
L-AZ91D-488–60– 488 60 1.501 8.7 0.20
1200G
L-AZ91D-600–60– 600 60 1.451 7.3 0.10
1200G
L-AZ91D-488–90– 488 90 1.507 13.0 0.29
1200G
L-AZ91D-600–90– 600 90 1.478 85.5 1.95
1200G

producing an homogeneous and continuous modified layer [53]. very different from that of samples with much better corrosion be-
Moreover, because of the fine nature of the b-Mg17Al12, it can be haviour, so that the strong change in the corrosion behaviour must
considered a continuous network of b-Mg17Al12, which seems to arise from factors associated with the surface morphology such as:
act as an effective barrier to the progression of corrosive attack. (i) the existence of defects in the form of voids as a consequence of
The macroscopic analysis of the corroded surface of this LSM treat- the high laser energies used [54], (ii) increase in the roughness of
ment (Fig. 11) indicated a homogeneous degradation of the surface, the surface which implies a higher surface area exposed to the cor-
which was confirmed by the cross-section SEM micrograph (Fig. rosive medium [51], (iii) incorporation of some carbon from the
14). In addition, there was a thin layer on the corroded surface. graphite paint used in the LSM treatments, and (iv) residual stress
As a consequence, the corrosion progression in this LSM treatment due to the high cooling rates [52].
is expected to start as a homogeneous dissolution of the a-Mg ma- Macroscopic views of LSM treated specimens (Figs. 3a, 4a and
trix enriched with aluminium, forming a thin layer on the treated 5a) revealed that surface roughness, surface voids and incorpora-
surface composed of magnesium-aluminium-oxygen (Fig. 14c and tion of graphite, increased with increasing laser input energy in
d) [24]. In addition, the progression of this corrosion mechanism the LSM treatments. This relationship was in agreement with the
was limited by the existing network of fine b-Mg17Al12, causing corrosion rates obtained by hydrogen evolution (Fig. 7) and elec-
the low corrosion rate. Nevertheless, the thin layer contained some trochemical measurements (Fig. 9), which again indicated that
defects (Fig. 14b) and the network of fine b-Mg17Al12 was not these factors had more influence on the corrosion behaviour of
completely continuous, so its effectiveness as a corrosion barrier the laser-treated specimens than the microstructure modifications
was limited. [51]. In fact, the cross-section analysis of the LSM treatment devel-
For the high input energy LSM treatments, the corrosion rates oped at 600 W and 60 mm s1 (Fig. 15) indicates that the corrosion
obtained from hydrogen evolution (Fig. 7) and polarization curves attack occurred because of voids in the surface of the treated sur-
(Fig. 9), indicated degradation in the corrosion behaviour of the face, which promoted crevice corrosion of the surface caused by
LSM treated samples. The microstructure in these cases were not micro-cells formed between voids (anodic behaviour) and the rest
C. Taltavull et al. / Materials and Design 57 (2014) 40–50 49

of the surface (cathodic behaviour) leading to the formation of pits 4) The LSM treatments caused an increase in the corrosion rate
inside the voids. Moreover, the small size of the voids intensified of the LSM, but after eliminating the top part of the treated
these phenomena. Once the pitting corrosion started, the micro- surface by grinding, the corrosion rate decreased up to 97%.
structure formed by the small isolated b-Mg17Al12 phase in a These results reveal that the effect of surface deformation,
a-Mg matrix (Fig. 5c) seemed to prevent the evolution of the residual stress, incorporation of carbon in the modified layer,
corrosion attack, but pits progressed through small heterogeneities and the production of voids in the surface, had more influ-
in the microstructure of the LSM treatment (Fig. 15b). Then corro- ence on the corrosion of the LSM treated samples than the
sion was controlled by the micro-galvanic couple between the microstructural modifications achieved.
b-Mg17Al12 phase and the a-Mg matrix, as in the as-received 5) After grinding, the corrosion rate of the samples was up to
AZ91D, and by the shape of the network of the b-Mg17Al12 phase 70% lower than those of the untreated AZ91D. This indicates
[1,2,5]. that the modification of the microstructure has a positive
To differentiate the role played by the microstructure and the effect in the corrosion evolution of the LSM samples.
surface defects, some samples were grounded and tested, so that 6) In general, laser treatment of AZ91D is a useful method to
the outer zone of the treated surface was removed. Therefore, the improve the corrosion resistance of the AZ91D alloy.
corrosion behaviour of these samples was only conditioned by its
microstructure. Unfortunately, this approach could not be made
for the SLSM samples, as the treated zone was small. Acknowledgements
There was a strong displacement of the polarization curves of
the LSM treated samples after being ground (Fig. 16). The corrosion Authors wish to thank Ministerio de Economia y Competitivi-
rates evaluated for the different samples (Fig. 17) confirmed this dad (project MAT2012–38407-C03–01) and associated FPI grant,
evolution, with corrosion rates that were in some cases two orders and to the Comunidad de Madrid (project S-0505/MAT/0077) for
of magnitude lower than in the as-treated state. Therefore, most of the funding, and the Australian Research Council Centre of
the corrosion observed in the laser treated samples was due to Excellence Design of Light Alloys for support.
surface defects, so the microstructures developed have a much
better corrosion resistance than the as-received samples [22–26].
References
The data presented in Fig. 17 and Table 3, indicate that the high-
est difference in the corrosion rates occurred in the LSM treated [1] Song G, Atrens. Understanding magnesium corrosion mechanism: a
samples at the highest input energies, whilst the lowest difference framework for improved alloy performance. Adv Eng Mater 2003;5:837.
[2] Atrens A, Liu M, Zainal Abidin NI. Corrosion mechanism applicable to
of corrosion rates occurred in the LSM treated samples at the
biodegradable magnesium implants. Mater Sci Eng B 2011;176:1609–36.
medium input energies. However, there was a wide range of para- [3] Walker J, Shadanbaz S, Kirkland NT, Stace E, Woodfield T, Staiger MP, et al.
meters for which the corrosion rate after surface grinding was well Magnesium alloys: predicting in vivo corrosion with in vitro immersion
below that of the as-received AZ91D alloy. In particular, for 600 W testing. J Biomed Mater Res B 2012;100B:1134–41.
[4] Abidin NI Zainal, Rolfe B, Owen H, Malisano J, Martin D, Hofstetter J, et al. The
and 60 mm s1 the corrosion rate was 79% lower than for the as- in vivo and in vitro corrosion of high-purity magnesium and magnesium alloys
received AZ91 Mg alloy. The fact that the corrosion rate after WZ21 and AZ91. Corros Sci 2013;75:354–66.
grinding was lower than that of AZ91D provides comfort that the [5] Song G, Atrens A. Corrosion mechanisms of magnesium alloys. Adv Eng Mater
1999;1:11–33.
grinding did not remove the laser surface treated material, because [6] Winzer N, Atrens A, Song G, Ghali E, Dietzel W, Kainer KU, et al. A critical
if the laser surface treated material had been removed leaving only review of the stress corrosion cracking (SCC) of magnesium alloys. Adv Eng
the AZ91D base material, the corrosion rate would have reflected Mater 2005;7:659–93.
[7] Atrens A, Song GL, Cao F, Shi Z, Bowen PK. Advances in Mg corrosion and
that of the AZ91D base material. research suggestions, Journal of Magnesium and Alloys, 2013;1: 177.
These results were in agreement with the fact that, as the laser [8] Taltavull C, Rodrigo P, Torres B, López AJ, Rams J. Dry sliding wear behavior of
input energy used increased, there was an increase in defects AM50B magnesium alloy. Materials and Design 2014;55:549–56.
[9] Mathieu S, Rapin C, Steinmetz J, Steinmetz P. A corrosion study of the main
introduced by LSM treatments, which degraded the corrosion constituent phases of AZ91 magnesium alloy. Corros Sci 2003;45:2741–55.
behaviour of the treated surface [54]. But also evidence that the [10] Song Y, Shan D, Chen R, Han EH. Effect of second phases on the corrosion
corrosion resistance increased by the elimination of big Mg17Al12 behaviour of wrought Mg–Zn–Y–Zr alloy. Corros Sci 2010;52:1830–7.
[11] Zhao MC, Liu M, Song G, Atrens A. Influence of the b-phase morphology on the
intermetallic precipitates and by the development of a finer
corrosion of the Mg alloy AZ91. Corros Sci 2008;50:1939–53.
b-Mg17Al12 phase network [17,22–26]. These observations were [12] Song G, Mishra R, Xu Q. Crystallographic orientation and electrochemical
in agreement with the prior corrosion literature [1,2,5,11,20,21]. activity of AZ31 Mg alloy. Electrochem Commun 2010;12:1009–12.
[13] Coy AE, Viejo F, Skeldon P, Thompson GE. Susceptibility of rare-earth-
magnesium alloys to micro-galvanic corrosion. Corros Sci 2010;52:3896–906.
5. Conclusions [14] Neil WC, Forsyth M, Howlett PC, Hutchinson CR, Hinton BRW. Corrosion of
magnesium alloy ZE41 – the role of microstructural features. Corros Sci
2009;51:387–94.
This research studied the influence of processing parameters on [15] Liu M, Uggowitzer PJ, Nagasekhar AV, Schmutz P, Easton M, Song G, et al.
the corrosion behaviour of laser surface modified AZ91D using a Calculated phase diagrams and the corrosion of die-cast Mg–Al alloys. Corros
high power diode laser. Sci 2009;51:602–19.
[16] Liu M, Qiu D, Zhao MC, Song G, Atrens A. The effect of crystallographic
orientation on the active corrosion of pure magnesium. Scripta Mater 2008;58:
1) A relationship was established between laser parameters 421–4.
used in the LSM treatment and the corrosion rate calculated [17] Guan YC, Zhou W, Li ZL, Zheng HY. Influence of overlapping tracks on
from both hydrogen evolution or electrochemical test: the microstructure evolution and corrosion behavior in laser-melt magnesium
alloy. Materials & Design 2013;52:452–8.
corrosion rate increased with increasing laser beam power
[18] Feng A, Han Y. Mechanical and in vitro degradation behavior of ultrafine
and decreasing laser scanning speed. calcium polyphosphate reinforced magnesium-alloy composites. Materials &
2) There was homogeneous corrosion attack and the formation Design 2011;32:2813–20.
of a thin surface layer for middle input energy LSM treat- [19] Wang L, Zhang BP, Shinohara T. Corrosion behavior of AZ91 magnesium alloy
in dilute NaCl solutions. wMaterials & Design 2010;31:857–63.
ments, implying a modification of the corrosion morphology [20] Cao F, Shi Z, Song GL, Liu M, Atrens A. Corrosion behaviour in salt spray and in
of the as-received AZ91D. 3.5 % NaCl solution saturated with Mg(OH)2 of as-cast and solution heat-
3) The SLSM treatments gave rise to corrosion rates that were treated binary Mg-X alloys: X = Mn, Sn, Ca, Zn, Al, Zr, Si, Sr,. Corrosion &
Science, 2013;76:60–97.
lower than the average measured for the AZ91D, especially [21] Shi Z, Cao F, Song GL, Liu M, Atrens A. Corrosion behaviour in salt spray and in
for the longest times tested. 3.5 % NaCl solution saturated with Mg(OH)2 of as-cast and solution heat-
50 C. Taltavull et al. / Materials and Design 57 (2014) 40–50

treated binary Mg-RE alloys: RE = Ce, La, Nd, Y, Gd,. Corrosion & Science, [38] Xu WL, Yue TM, Man HC, Chan CP. Laser surface melting of aluminium alloy
2013;76:98–118. 6013 for improving pitting corrosion fatigue resistance. Surf Coat Technol
[22] Mondal AK, Kumar S, Blawert C, Dahotre NB. Effect of laser surface treatment 2006;200:5077–86.
on corrosion and wear resistance of ACM720 Mg alloy. Surf Coat Technol [39] Viejo F, Coy AE, Garcia-Garcia FJ, Liu Z, Skeldon P, Thompson GE. Relationship
2008;202:3187–98. between microstructure and corrosion performance of AA2050-T8 aluminium
[23] Khalfaoui W, Valerio E, Masse JE, Autric M. Excimer laser treatment of ZE41 alloy after excimer laser surface melting. Corros Sci 2010;52:2179–87.
magnesium alloy for corrosion resistance and microhardness improvement. [40] Rams J, Pardo A, Urena A, Arrabal R, Viejo F, Lopez AJ. Surface treatment of
Opt Laser Eng 2010;48:926–31. aluminium matrix composites using a high power diode laser. Surf Coat
[24] Coy AE, Viejo F, Garcia-Garcia FJ, Liu Z, Skeldon P, Thompson GE. Effect of Technol 2007;202:1199–203.
excimer laser surface melting on the microstructure and corrosion [41] Viejo F, Pardo A, Rams J, Merino MC, Coy AE, Arrabal R, et al. High power diode
performance of the die cast AZ91D magnesium alloy. Corros Sci 2010;52: laser treatments for improving corrosion resistance of A380/SiCp aluminium
387–97. composites. Surf Coat Technol 2008;202:4291–301.
[25] Abbas G, Liu Z, Skeldon P. Corrosion behaviour of laser-melted magnesium [42] Liu Z, Chong PH, Skeldon P, Hilton PA, Spencer JT, Quayle B. Fundamental
alloys. Appl Surf Sci 2005;247:347–53. understanding of the corrosion performance of laser-melted metallic alloys.
[26] Banerjee PC, Raman RK Singh, Durandet Y, McAdam G. Electrochemical Surf Coat Technol 2006;200:5514–25.
investigation of the influence of laser surface melting on the microstructure [43] Taltavull C, Torres B, Lopez AJ, Rodrigo P, Otero E, Rams J. Selective laser surface
and corrosion behaviour of ZE41 magnesium alloy-An EIS based study. Corros melting of a magnesium–aluminium alloy. Mater Lett 2012;85:98–101.
Sci 2011;53:1505–14. [44] Lin L. The advances and characteristics of high-power diode laser materials
[27] Abbas G, Li L, Ghazanfar U, Liu Z. Effect of high power diode laser surface processing. Opt Laser Eng 2000;34:231–53.
melting on wear resistance of magnesium alloys. Wear 2006;260:175–80. [45] Kirkland NT, Birbilis N, Staiger MP. Assessing the corrosion of biodegradable
[28] Zhang Y, Chen J, Lei W, Xv R. Effect of laser surface melting on friction and wear magnesium implants: a critical review of current methodologies and their
behaviour of AM50 magnesium alloy. Surf Coat Technol 2008;202:3175–9. limitations. Acta Biomater 2012;8:925–36.
[29] Lv XX, Liu HY, Wang YB, Lu Y, Li, An. Microstructure and dry sliding wear [46] Song GL, Atrens A, StJohn D. An hydrogen evolution method for the estimation
behavior of Mg–Y–Zn alloy modified by laser surface melting. J Mater Eng of the corrosion rate of magnesium alloys. In: Hryn JN, editor. Magnesium
Perform 2011;20:1015–22. technology 2001 symposium, minerals. New Orleans, LA: Metals & Materials
[30] Kwok CT, Lo KH, Chan WK, Cheng FT, Man HC. Effect of laser surface melting on Soc; 2011. p. 255–62.
intergranular corrosion behaviour of aged austenitic and duplex stainless [47] Shi Z, Prasad A, Atrens A, Plug-in specimens for the measurement of the
steels. Corros Sci 2011;53:1581–91. corrosion rate of Mg alloys JOM 2012;64(6):657–63.
[31] Kwok CT, Lo KH, Cheng FT, Man HC. Effect of processing conditions on the [48] Shi Z, Atrens A. An innovative specimen configuration for the study of Mg
corrosion performance of laser surface-melted AISI 440C martensitic stainless corrosion. Corros Sci 2011;53:226–46.
steel. Surf Coat Technol 2003;166:221–30. [49] Taltavull C, López AJ, Torres B, Rams J. Fracture behaviour of a magnesium-
[32] Kwok CT, Cheng FT, Man HC. Microstructure and corrosion behaviour of laser aluminium alloy treated by selective laser surface melting treatment.
surface-melted high-speed steels. Surf Coat Technol 2007;202:336–48. Materials and Design 2014;55:361–5.
[33] Bao G, Shinozaki K, Iguro S, Inkyo M, Yamamoto M, Mahara Y, et al. Stress [50] Shi Z, Jia JX, Atrens A. Galvanostatic anodic polarisation curves and galvanic
corrosion cracking sealing in overlaying of Inconel 182 by laser surface corrosion of high purity Mg in 3.5%NaCl saturated with Mg(OH)2. Corros Sci
melting. J Mater Eng Perform 2006;173:330–6. 2012;60:296–308.
[34] Li R, Ferreira MGS, Almeida A, Vilar R, Watkins KG, McMahon MA, et al. [51] Walter R, Kannan MB. Influence of surface roughness on the corrosion
Localized corrosion of laser surface melted 2024-T351 aluminium alloy. Surf behaviour of magnesium alloy. Materials and Design 2011;32:2350–4.
Coat Technol 1996;81:290–6. [52] Kouadri A, Barrallier L. Study of mechanical properties of AZ91 magnesium
[35] Chong PH, Liu Z, Skeldon P, Thompson GE. Large area laser surface treatment of alloy welded by laser process taking into account the anisotropy
aluminium alloys for pitting corrosion protection. Appl Surf Sci 2003;208– microhardness and residual stresses by X-ray diffraction. Metall Mater Trans
209:399–404. A 2011;42A:1815–26.
[36] Osorio WR, Cheung N, Spinelli JE, Cruz KS, Garcia A. Microstructural [53] Taltavull C, Torres B, López AJ, Rodrigo P, Rams J. Novel laser surface
modification by laser surface remelting and its effect on the corrosion treatments on AZ91 magnesium alloy. Surface and Coatings Technology
resistance of an Al–9 wt%Si casting alloy. Appl Surf Sci 2008;254:2763–70. 2013;222:118–27.
[37] Padovani C, Davenport AJ, Connolly BJ, Williams SW, Siggs E, Groso A, et al. [54] Laazizi A, Courant B, Jacquemin F, Andrzejewski H. Applied multi-pulsed laser
Corrosion protection of AA7449-T7951 friction stir welds by laser surface in surface treatment and numerical-experimental analysis. Optics & Laser
melting with an Excimer laser. Corros Sci 2011;53:3956–69. Technology 2011;43:1257–63.

You might also like