Download as pdf or txt
Download as pdf or txt
You are on page 1of 16

Applied Physics A (2021) 127:186

https://doi.org/10.1007/s00339-021-04331-0

Effects of Ag content and heat treatment on the microstructure


and properties of SLMed AZ61 Mg–Al–Zn alloy
Chenguang Hu1 · Changjun Chen1 · Min Zhang1

Received: 20 December 2020 / Accepted: 25 January 2021 / Published online: 16 February 2021
© The Author(s), under exclusive licence to Springer-Verlag GmbH, DE part of Springer Nature 2021

Abstract
The effects of Ag (0, 0.5, 1, and 2 wt%) content on the microstructure and mechanical properties of SLMed AZ61 Mg–Al–Zn
alloy in the as-deposited and heat-treated conditions were investigated. The microstructure of SLMed AZ61 Mg–Al–Zn alloy
is composed of α-Mg matrix and eutectic of α-Mg and M ­ g17Al12 phase at grain boundaries. The addition of Ag inhibited the
growth of β-Mg17Al12 phase and further refined the grain size. After adding 2 wt% Ag, the volume fraction of β-Mg17Al12
phase is decreased from 40.53 to 30.27%, and the average grain diameter is decreased from 2.35 to 1.82 μm. ­Mg54Ag17 phase
particles were formed after Ag was added and distributed in the grain boundary, which was beneficial to enhance the mechani-
cal properties of Mg–Al alloy. After 8-h solution treatment at 420 °C β-Mg17Al12 phase was basically dissolved in α-Mg
matrix, which was beneficial to the precipitation of discontinuous β phase during aging treatment, and this discontinuous β
phase could effectively enhance the strength of Mg–Al alloy. When the content of Ag is 2 wt%, the mechanical properties and
corrosion resistance of Mg–Al alloy are enhanced. After solution treatment, the mechanical properties of Mg–Al alloy are
decreased, but the corrosion resistance is greatly enhanced. After aging treatment, the mechanical properties are increased,
but the corrosion resistance is decreased.

Keywords Mg–Al–Zn alloy · Ag addition · Selective laser melting · Heat treatment · Microstructure

1 Introduction defects. In the process of plastic deformation, the plasticity


and strength of the material may decrease [3]. Magnesium
With the increasing demand for lighter weight and improved alloys made by powder metallurgy may not be dense enough
fuel economy in the aerospace and automotive industries, to be strong enough [4]. Selective laser melting (SLM) is
magnesium alloys have attracted widespread attention [1, an additive manufacturing process. The laser runs along a
2]. However, the application of magnesium alloy is limited specific path, and the metal powders are deposited layer by
because of its low strength. At present, the main manufac- layer on the experimental platform [5]. Compared with other
turing method of magnesium alloy is casting. In addition, manufacturing methods, SLM can process the parts with
plastic deformation and powder metallurgy are also used in more complex geometry [6]. SLM can also be used to make
the preparation of magnesium alloys. However, the above magnesium alloy structural parts. At present, there are some
methods all have their own disadvantages, for casting, easy researches about selective laser melting to make magnesium
to produce shrinkage and porosity and inclusion and other alloy. It is well known that the reduction of grain size has
a great effect on the mechanical properties of metals, espe-
cially magnesium alloys. In Hall–Petch relationship, it has
* Changjun Chen
503047820@qq.com been proved that the strengthening coefficient of magnesium
alloy is four times than that of aluminum alloy [7]. If SLM
Chenguang Hu
1475198348@qq.com is applied to manufacture parts of magnesium alloy, the
characteristics of rapid cooling can overcome the defects of
Min Zhang
63606536@qq.com micro- and macro-component segregation and coarse grains
in the manufacturing process and finally obtain magnesium
1
Laser Processing Research Center, School of Mechanical alloys with good properties [8].
and Electric Engineering, Soochow University,
Suzhou 215131, China

13
Vol.:(0123456789)
186 Page 2 of 16 C. Hu et al.

However, the addition of trace elements can further the powder morphology after mixing the two are shown in
improve the mechanical properties and corrosion resist- Fig. 1. The morphology of the Mg–Al alloy powder used in
ance of laser additive manufacturing of magnesium alloys. this study is shown in Fig. 1c, showing a regular spherical
In recent years, in order to improve the mechanical prop- shape with good fluidity, and the powder diameter is about
erties of magnesium alloys, Ca, Ti, Sr, and rare earth ele- 20–80 μm. The diameter of Ag powder is about 10-50 μm,
ments have been widely studied and applied [9]. Many stud- as shown in Fig. 1b. After the two powders are uniformly
ies have shown that the addition of trace elements to AZ mixed, the powder morphology is shown in Fig. 1a.
series magnesium alloys is an effective way to improve and All the samples in this experiment were prepared on the
optimize the mechanical properties of Mg–Al alloys. The TH-LWY300 high-speed laser system produced by Suzhou
microstructure and mechanical properties of AZ series mag- Tianhong Laser Co., Ltd. The system mainly includes YAG-
nesium alloys can be improved by adding trace Ag element 300 W pulsed laser, computer control system, three-degree-
[10]. Ag can effectively improve the strength of magnesium of-freedom workbench, and argon gas protection system.
alloy [11]. ­Mg17Al12 phase with body-centered cubic struc- Before the start of the experiment, the cast Mg–Al alloy sub-
ture is formed in Mg–Al alloys during cooling from high strate (AZ61) was fixed on the workbench, and a marker is
temperature to low temperature, and ­Mg17Al12 phase plays used to blacken the substrate to increase the absorption rate
an important role in the properties of AZ series magnesium of the laser during sample preparation, because the substrate
alloys. The mechanical properties of Mg–Al alloys are poor, has a higher reflectivity and the focal length of the pulsed
especially at high temperature, which is mainly due to the laser system is 136 mm. Then the experimental powder is
low hardness of ­Mg17Al12 phase at grain boundary. Adding spread evenly on the substrate, the powder under the laser
a small amount of Ag in AZ series magnesium alloy can scanning path is melted, then the next laser scanning path
improve the density of M ­ g17Al12 phase and the density of rotates 90 degrees, and the two laser scanning paths alter-
­Mg17Al12 phase, so as to increase the strength of magnesium nately run during the experiment. The laser parameters used
alloy [12]. Krzysztof Bryla et al. found that after adding 4 during sample preparation are listed in Table 2.
wt% Ag to the casting Mg, a M ­ g54Ag17 strengthening phase The heat treatment experiment in this experiment is car-
was formed, and after the T4 heat treatment, the mechani- ried out in a high-temperature vacuum tube furnace (GSL-
cal properties were greatly improved [13]. Jiashi Miao et al. 1700X-80X) to avoid oxidation of magnesium alloy at high
added 0.7 wt% Ag to the Mg–Al–Sn alloy to increase the temperature. There are two heat treatment methods for the
aging hardening effect of the magnesium–aluminum alloy, sample: solution; solution and aging treatment. And the
thereby increasing the strength, ductility, and corrosion influence of different solution time on magnesium alloy was
resistance of the Mg–Al–Sn alloy [13]. However, there are explored in this study, The solution temperature is 420 °C,
few researches about the effect of adding Ag on the laser and the solution time is 6 h and 8 h, respectively. The tem-
additive manufacturing of magnesium alloys (Mg–Al–Zn). perature of aging experiment is 200 °C, and the aging time
This article aims to study the structure and properties of is 10 h. The heat treatment temperature rise curve is shown
laser additive manufacturing of magnesium alloys, and in Fig. 2.
on this basis, by adding Ag and heat treatment to further Before microstructure observation, all of the samples
improve the performance of magnesium aluminum alloy. (deposited samples and heat-treated samples) were sub-
jected to grind and polish with sandpaper and polishing
paste of different grand sizes. In this experiment, sand-
2 Experimental procedures paper with grand sizes of 600, 1000, and 2000 was used
to polish the sample on a polishing machine (model:
Ag powder of different proportions was mixed into atom- MP-2A). Then water-soluble diamond polishing paste with
ized AZ61 powder by using ball grinder for 8 h. The powder a particle size of 1.0 W was used for polishing. After pol-
composition used in the experiment is shown in Table 1. ishing, the sample is placed in alcohol and cleaned in an
Magnesium aluminum alloy powder and Ag powder and ultrasonic cleaning machine (HQD-200). Before observing

Table 1  Chemical compositions Experimental alloys Powder composition (wt%)


of experimental alloys (wt%)
Al Zn Ag Mn Si Mg

AZ61 6.08 1 – 0.23 0.038 Bal.


AZ61 + 0.5%Ag 6.08 1 0.5 0.23 0.038 Bal.
AZ61 + 1%Ag 6.08 1 1 0.23 0.038 Bal.
AZ61 + 2%Ag 6.08 1 2 0.23 0.038 Bal.

13
Effects of Ag content and heat treatment on the microstructure and properties of SLMed AZ61 Mg–… Page 3 of 16 186

Fig. 1  a AZ61 + 2%Ag powder,


b Ag powder, c AZ61 powder

Table 2  SLM manufacturing parameters the microstructure, the polished and cleaned sample is
etched in the etching solution. The etching solution in this
Processing parameters Value
experiment is: 4% nitric acid + 96% alcohol. Etching time
Laser power P, W 90 is 60–70 s. The microstructure and element distribution
Pulse width t, ms 2.5 scanning (EDS) experiment of the sample was observed
Frequency f, Hz 40 and photographed under Zeiss EVO 18 scanning electron
Hatch spacing l, mm 0.4 microscope, the intersection method was used to calculate
Scanning velocity v, mm/s 10 the Mg–Al alloy grain size, and the image processing soft-
Layer thickness T, mm 0.2 ware was used to calculate the ratio of ­Mg17Al12 phase. In
Light spot diameter d, mm 0.1 the experiment, an X-ray diffractometer (model: X’-Pert-
Pro MPD) was used to analyze the phase of samples, a
micro-hardness tester (model: HXD-1000TMC/LCD) was
used to test the micro-hardness of the samples, the micro-
hardness test load is 100 g, and the loading time is 10 s.
The hardness value of the sample is the average of 10 ran-
dom measurements. The micro-scratching and nano-inden-
tation experiments were performed on the CSM micro-
scratching tester. The micro-scratching load was 100 mN
or 300 mN, and the scratch length was 500 μm. The load
of the nano-indentation is 5 mN, and the nano-hardness
value is the average of 6 measurements. CHI700E elec-
trochemical workstation was used to measure the polari-
zation curve. The working electrode, reference electrode,
and auxiliary electrode of the electrochemical workstation
are, respectively, the sample to be tested with an effective
area of 0.4 ­cm2, a saturated calomel electrode (SCE), and
a platinum sheet. After polishing the sample, put it in 3.5%
NaCl corrosive medium and let it stand for 60 min. After
the open circuit potential is stable, the measurement is
Fig. 2  Heat treatment process carried out at room temperature, and the scanning speed

13
186 Page 4 of 16 C. Hu et al.

is 1 mV/s. Then chi750e software is used to calculate the Table 3  The grain size and β-Mg17Al12 of Mg–Al alloy with different
self-corrosion potential and self-corrosion current density. proportions of Ag
Alloy AZ61 AZ61 + 0.5Ag AZ61 + 1Ag AZ61 + 2Ag

Grain size/μm 2.35 2.17 2.25 1.82


3 Results and discussion
Volume 40.53% 37.01% 33.69 30.27%
fraction of
3.1 Effect of Ag on the structure and properties ­Mg17Al12
of Mg–Al–Zn–Ag alloy

3.1.1 Microstructure of as‑deposited AZ61 Mg alloys than that of the cast magnesium alloy, and the formation of
β-Mg17Al12 phase is related to many factors, and the cool-
The microstructure of SLMed Mg–Al alloy is shown in ing rate is a very important factor, the high cooling rate can
Fig. 3, which is mainly composed of α-Mg matrix and promote the formation of β-Mg17Al12 phase. Therefore, the
β-Mg17Al12 phase formed during non-equilibrium solidifi- phase ratio of β-Mg17Al12 in SLMed Mg–Al alloy is much
cation. After adding different proportions of Ag, the grain higher than that of cast magnesium alloy. The grain size will
size of Mg–Al alloy did not change significantly. With the affect the mechanical properties such as the hardness of the
increase in Ag content, the proportion of β-Mg17Al12 phase material. In the AZ magnesium alloy, the existence form and
gradually decreased, as shown in Fig. 3b, c; the original proportion of the β-Mg17Al12 phase determine the properties
continuous network eutectic phase is broken; in Fig. 3d, of the magnesium alloy. These characteristics determine that
the thickness of the β-Mg17Al12 phase located at the grain the mechanical properties and corrosion resistance of laser
boundary is significantly reduced. The grain size and the additive manufacturing of magnesium alloys are very differ-
volume fraction change of β-Mg17Al12 phase are listed in ent from those of cast magnesium alloys.
Table 3. The decrease in β-Mg17Al12 phase indicates that the Figure 4 shows the microstructure of SLMed Mg–Al
addition of Ag may change the non-equilibrium solidifica- alloy. It can be seen that the eutectic phase at the grain
tion process of Mg–Al alloy and delay the eutectic reaction. boundary is composed of α-Mg and β-Mg17Al12. In Mg–Al
Due to the characteristics of rapid melting and rapid cool- alloys, the morphology of the eutectic phase has two forms:
ing in the laser additive manufacturing process, the micro- partially divorced eutectic and completely divorced eutectic.
structure of laser additive manufacturing of magnesium Factors such as the cooling rate, the aggregation of solid
alloys is very different from that of cast magnesium alloys. solution atoms, and the morphology of the crystal grains
For example, in the short cooling process, the grains have will affect the morphology of the eutectic phase [14]. During
no time to grow further, so the grain size is much smaller the cooling and solidification process, first α-Mg dendrites

Fig. 3  SEM micrograph of


as-deposited AZ61 + xAg: a
x = 0%, b x = 0.5%, c x = 1%, d
x = 2%

13
Effects of Ag content and heat treatment on the microstructure and properties of SLMed AZ61 Mg–… Page 5 of 16 186

of the β-Mg17Al12 phase and precipitation, the Mg atoms


are extruded side by side, and the α-Mg nucleation condi-
tions are formed after the Mg atoms aggregate. Al atoms
are ejected to the grain boundaries due to the nucleation of
α-Mg [15]. In the process of laser additive manufacturing,
the eutectic phase also grows in this way alternately. But
there is much less time for α-Mg and β-Mg17Al12 phase to
grow because of the high cooling rate. The partially divorced
eutectic morphology is shown in Fig. 4b.
The XRD pattern of the unheated Mg–Al alloy is shown
in Fig. 5. The phases of all samples are mainly composed of
α-Mg phase and β-Mg17Al12 phase. According to the EDS
element distribution map and XRD pattern, after adding Ag
element, a new phase M ­ g54Ag17 is formed, the correspond-
ing peak is near the main peak, and with the increase in
Ag content, the corresponding peak height of M ­ g54Ag17
increases significantly. The ε′-Mg54Ag17 phase has a body-
centered orthorhombic structure. According to the Mg-Ag
binary phase diagram, the ε′-Mg54Ag17 phase is a high-tem-
perature resistant phase formed in the peritectic reaction or
the eutectic reaction, and it is also a kind of strengthening
phase in the magnesium alloy [16–18].
As shown in Fig. 6, the content of Mg element at the
grain boundary is lower than the content of Mg in the α-Mg
matrix, while the content of Al element is on the contrary. Al
element aggregates at the grain boundary, and the content of
Al element is higher than the Al content in the matrix. Com-
Fig. 4  Microstructure and morphology of deposited Mg–Al–Zn– bined with the XRD pattern and the Mg–Al alloy micro-
0.5Ag alloy samples
structure, the grain boundary is an eutectic of α-Mg and
β-Mg17Al12. However, the Ag element exists in the α-Mg
begin to nucleate, and the solute atoms are pushed to the matrix and the grain boundary in the SLMed Mg–Al alloy,
solid–liquid interface, so the concentration of solute atoms and the Ag content at the grain boundary is significantly
at the solid–liquid interface is getting higher and higher. higher than the Ag content in the matrix. This shows that
In the eutectic reaction, due to the higher melting point the additive Ag is segregated in β-Mg17Al12 phase due to

Fig. 5  The XRD pattern of the


unheated Mg–Al–Zn–Ag alloy

13
186 Page 6 of 16 C. Hu et al.

Fig. 6  Microstructure and element distribution of deposited Mg–Al–Zn–2Ag alloy

its good diffusibility and chemical affinity with Al. Accord-


ing to the qualitative thermodynamic standard proposed by
Mendis et al., precipitation (second phase) can be strength-
ened by adding a small amount of alloying elements [19].
According to the XRD pattern, with the increase in Ag con-
tent, the peak corresponding to ­Mg54Ag17 is detected more
and more obvious. Combined with the distribution of Ag
element, ­Mg54Ag17 particles should exist in the β-Mg17Al12
phase. The Zn element in the alloy promotes the forma-
tion of the M­ g54Ag17 phase. The crystal structure of the
­Mg51Zn20 phase and the ­Mg54Ag17 phase is basically the
same, and because the diffusion coefficient of Ag in the
α-Mg matrix is much greater than that of Zn, this means that
during the cooling and crystallization process, the Ag atom
may replace the position of the Zn atom to form a replace-
ment solid solution [20, 21]. Therefore, Zn atoms should be
dispersed in the M­ g54Ag17 phase, and the M ­ g54Ag17 phase
mainly exists in the β-Mg17Al12 phase at the grain boundary, Fig. 7  Micro-hardness of Mg–Al–Zn–Ag alloy samples with different
so the content of Zn element at the grain boundary should Ag contents
be higher than the matrix, as shown in Fig. 6 Zn element
distribution diagram.
Mg–Al alloys is much larger than casting magnesium alloy,
3.1.2 Mechanical properties of as‑deposited AZ61 Mg which is also an important reason for the increase in hard-
alloys ness. It can be seen that when the content of Ag is 0.5%
and 1%, the hardness value increases slightly to 91.3HV
The micro-hardness of the as-deposited Mg–Al alloy sample and 93.8HV, respectively. However, when the Ag content
is shown in Fig. 7. Before heat treatment, the hardness value is increased to 2%, the hardness value is the highest, reach-
of as-deposited Mg–Al alloy is between 86.7 and 109.0HV. ing 105.3HV, which is 21.45% higher than the hardness of
Compared with the micro-hardness of 58 HV of traditional the sample without Ag. As discussed before, the addition
cast magnesium alloy [22], the hardness of laser additive of Ag will further refine the grains of the SLMed Mg–Al
manufacturing magnesium alloy is significantly improved. alloy; although the addition of Ag will reduce the propor-
The grain size of the magnesium alloy produced by SLM is tion of β-Mg17Al12 phase, Ag can increase the density of
much smaller than that of the cast magnesium alloy. Accord- β-Mg17Al12 phase to improve the mechanical properties of
ing to the Hall–Petch theory, the reduction of the grain size Mg–Al alloy [23]. Through the element distribution diagram
is beneficial to increase the hardness of the material; in of Ag, it can be seen that part of the Ag element is solid-dis-
addition, the proportion of the β-Mg17Al12 phase of SLMed solved in the α-Mg matrix; these will increase the hardness

13
Effects of Ag content and heat treatment on the microstructure and properties of SLMed AZ61 Mg–… Page 7 of 16 186

of the SLM magnesium alloy. In addition, the M ­ g54Ag17 Table 4  Young’s modulus and indentation depth of samples with dif-
phase distributed at the grain boundary can also strengthen ferent Ag contents under 5 mN load
the magnesium alloy. Ag HIT (MPa) EIT (GPa) Penetration depth/ Rebound
The results of the nano-indentation experiment of Mg–Al content residual depth (nm) rate (%)
alloy with different proportions of Ag content are shown in (wt%)
Fig. 8 and Table 4. The loading curve of the nano-indenta- 0 1351.4 39.81 322/228 70.8
tion test load is shown in Fig. 8a, linear loading from 0 to 0.5 1500.3 44.17 302/209 69.2
5 mN, holding the load for 10 s, and then linear unloading. 1 1693.7 46.85 292/213 72.9
The measured nano-hardness trend is consistent with the 2 2224.3 48.62 264/192 72.7
micro-hardness, but the nano-hardness excludes the compre-
hensive influence of pores and other defects, and the indenta-
tion area is small, so the grain size has less influence on the analysis. As shown in Fig. 8c, all samples have different
nano-hardness. The nano-hardness of the sample without degrees of yield deformation, and the sample without Ag
Ag element is 1351.4 MPa. When the Ag content is 0.5% element has the largest yield deformation. With the increase
and 1%, the nano-hardness increases slightly to 1500.3 MPa in Ag content, the maximum depth of indentation and resid-
and 1693.7 MPa, respectively. When the Ag content is 2%, ual depth both decrease. When the Ag content is 2%, the
the nano-hardness is the highest, reaching 2224.3 MPa. The maximum depth of the indentation and the residual depth
nano-hardness mainly depends on the enhancement of the are the minimum: 268/188 nm. It is shown that after Ag ele-
β-Mg17Al12 phase by the Ag element and the solid solu- ment is added, the rigidity of magnesium aluminum alloy is
tion strengthening effect of the Ag element in the matrix, enhanced. It can be seen from Table 4 that the resilience rate
which is consistent with the micro-hardness and structure and the elastic modulus basically show a linear relationship,

Fig. 8  Nanoindentation experiment results of samples with different Ag contents: nanoindentation load curve; b average nanohardness; c load
displacement curve; d indentation depth curve

13
186 Page 8 of 16 C. Hu et al.

the larger the elastic modulus, the greater the resilience


rate. After Ag is added, the elastic modulus of Mg–Al alloy
is increased to varying degrees, which may be due to the
change in the microstructure caused by the addition of Ag;
thereby, the Young’s modulus is increased.
Mg–Al–Zn–Ag alloy samples were carried out micro-
scratching experiment under 100 or 300 mN, and the
residual depth curve is shown in Fig. 9. When the experi-
mental load is 100mN, the residual depth of the 2%Ag
sample is between 400 and 600 nm, and the residual depth
is reduced by about 200 nm compared with the sample
without Ag. When the load is 300 mN, the same trend is
shown, and the residual depth curve is more smooth. The
average value of the residual depth of the 2% Ag sample
is 914 nm, the average value of the residual depth of the
sample without Ag element is 1901 nm, and the residual
depth is reduced by about 987 nm. It is shown that the
addition of Ag enhances the resistance of the sample to
micro-scratching. Figure 10 shows the comparison of pen-
etration depth and residual depth. In the sample without
Ag, the average penetration depth is 2503 nm, and the
ratio of the average residual depth to penetration depth is
75.9%. In the 2% Ag sample, the average penetration depth
Fig. 10  Penetration depth and residual depth of Mg–Al–Zn-xAg alloy
is 1781 nm, and the ratio of the average residual depth to sample under 300mN load a x = 0; b x = 2
the penetration depth is 51.3%. It is shown that the addi-
tion of Ag reduces the resilience rate, combined with the
elastic modulus value and micro-hardness value obtained from the nano-indentation experiment; after adding 2%
Ag to the SLMed Mg–Al alloy, the hardness of the mate-
rial increases, and the plasticity decreases. And through
the comparison of the residual depth and the penetration
depth value, it is found that the addition of Ag signifi-
cantly reduces the depth of micron scoring and effectively
improves the ability of Mg–Al alloy to resist external force
intrusion. The main reason for increasing the strength of
magnesium–aluminum alloy may be the strengthening of
the β-Mg17Al12 phase by the Ag element and the M ­ g54Ag17
strengthening phase distributed and the grain boundary,
which effectively prevents the intrusion of the micron
scratching probe.
Figure 11 shows the friction coefficient curve of the sam-
ple with 2% Ag and non-Ag through the micro-scratching
experiment. When the load is 100 mN, the average friction
coefficient of the sample with Ag content of 0% and 2%
is, respectively, 0.31 and 0.22. When the load increased
to 300 mN, the friction coefficients of the two groups of
samples increased. The average friction coefficients of the
samples with Ag content of 0% and 2% were 0.35 and 0.27,
respectively. The penetration depth increases after the load
increases, resulting in an overall increase in the friction
coefficient. However, under both loads, the addition of Ag
reduces the friction coefficient of the Mg–Al alloy. Accord-
Fig. 9  Residual depth of Mg–Al–Zn–Ag alloy sample under different ing to Archard’s law [24], the friction coefficient increases
load a 100 mN b 300 mN when the hardness decreases accordingly, which further

13
Effects of Ag content and heat treatment on the microstructure and properties of SLMed AZ61 Mg–… Page 9 of 16 186

Fig. 12  Polarization curves of samples with different Ag content

is more than100 mV, and the precipitation of intermetallic


compound is continuous, the intergranular corrosion of the
metal can occur [25]. For the laser additive manufacturing of
Fig. 11  Friction coefficient curve of Mg–Al–Zn–Ag alloy sample
under different load. a 100 mN; b 300 mN Mg–Al alloys, the network β-Mg17Al12 phase is distributed
between the grain boundaries, the self-corrosion potential
of the β-Mg17Al12 phase is lower than that of the Mg matrix
shows that the hardness of Mg–Al alloy increases after Ag [26], the network β phase tends to form electrochemical
is added, and the wear resistance improves. corrosion, and the corrosion resistance of Mg–Al alloy is
reduced. According to the above microstructure analysis,
3.1.3 Corrosion resistance of as‑deposited Mg–Al–Zn–Ag the addition of Ag inhibits the formation of β-Mg17Al12
alloy phase, so it is beneficial to improve the corrosion resistance
of Mg–Al alloy.
The Tafel curves measured by samples with different Ag Silver atoms have a relatively high standard potential
content are shown in Fig. 12. Using the Tafel curve extrap- (0.78 V), while the standard potential of magnesium matrix
olation method, the self-corrosion potential and corrosion is relatively low (− 1.67 V) [27]. Therefore, the Ag element
current density of the four groups of samples in 3.5% NaCl dissolved in the α-Mg can increase the potential of the α-Mg.
solution are calculated and listed in Table 5. The lowest self- In addition, the self-corrosion potential of the M ­ g54Ag17
corrosion potential of the sample without Ag is − 1.285 V, phase distributed and the grain boundary is higher than that
and the self-corrosion current density is 9.144 × 10 –4 of the β-Mg17Al12 phase, so the M ­ g54Ag17 phase near the
A cm−2. When the Ag content reaches 2%, the self-corrosion grain boundary can enhance the electrochemical corrosion
potential of the sample is the highest: − 1.262 V. Compared performance along the grain boundary, thereby increasing
with the sample without Ag, the self-corrosion potential is the corrosion resistance of Mg–Al alloy.
increased by 23 mV, and the self-corrosion current density
is also reduced to 6.656 × 10–4 A cm−2. It is shown that the
corrosion tendency and corrosion rate of the 2% Ag sam-
ple are lower than that of the sample without Ag. The self- Table 5  Self-corrosion potential and corrosion current density of
corrosion potential of samples with Ag content of 0.5% and samples with different Ag content
1% is between − 1.285 and − 1.262 V, indicating that the Samples Ecoor(vs. SCE)/V Jcoor/(A cm−2)
corrosion resistance is also improved.
AZ61 – 1.285 9.144 × 10–4
The corrosion resistance of Mg–Al alloys largely depends
AZ61 + 0.5%Ag – 1.279 7.260 × 10–4
on the morphology and distribution of β-Mg17Al12 phase.
AZ61 + 1%Ag – 1.277 6.848 × 10–4
When the corrosive medium exists, the potential difference
AZ61 + 2%Ag – 1.262 6.656 × 10–4
between the intermetallic compound and the solid solution

13
186 Page 10 of 16 C. Hu et al.

3.2 Effect of heat treatment on the structure alloy sample after solution treatment. Al element has obvi-
and properties of SLMed Mg–Al–Zn–Ag alloy ous massive aggregation, while the content of Mg element
is relatively reduced at the Al aggregation site, which proves
3.2.1 Microstructure of heat‑treated AZ61 Mg alloys that after 6 h of solution treatment, the white block phase
is β-Mg17Al12 phase. When the solution time is extended
The temperature of the solid solution experiment is 420 °C, to 8 h, the morphology of the Mg–Al alloy is shown in
and the holding time is 6 h or 8 h in order to study the effect Fig. 13e, f, and almost no block β phase is observed. As
of different solid solution time on the solid solution effect. the solution time increases, the β phase basically dissolved
Figure 13a, b shows the microstructure of the as-deposited into the Mg matrix, leaving only a few long strips of β
Mg–Al–Zn–2Ag sample, with eutectic phase at the grain phase. After T4 treatment, most of the netlike and strip-
boundary. After 6 h of solution treatment, the original net- like β-Mg17Al12 phases produced in the non-equilibrium
like β-Mg17Al12 phase and α-Mg eutectic phase were trans- solidification state have been dissolved in the matrix, and
formed into block and rodlike β phases [14]. Figure 14 shows a small part of the Al element is not completely dissolved
the element distribution diagram of the Mg–Al–Zn–2Ag in the matrix. The remained Al element is dispersed and

Fig. 13  Microstructure of Mg–Al–Zn–2Ag alloy under different heat treatment condition a, b as-deposited c, d solid solution for 6 h e, f solid
solution for 8 h

13
Effects of Ag content and heat treatment on the microstructure and properties of SLMed AZ61 Mg–… Page 11 of 16 186

Fig. 14  Element distribution of Mg–Al–Zn–2Ag alloy sample after solution treatment

distributed in the α-Mg matrix as fine granular β-Mg17Al12 grow and coarsen, which will adversely affect the mechani-
phases through diffusion effect. Similar to the solid solution cal properties of the material [29]. From the analysis of the
process of as-cast Mg–Al alloy, after 8 h of solution treat- XRD diffraction results of the Mg–Al–Zn–2Ag alloy after
ment, the Al element that is not dissolved in the α-Mg matrix heat treatment, after 6 h solution heat treatment, the corre-
is distributed in the matrix as a thin rod-shaped β-Mg17Al12 sponding peak of M ­ g54Ag17 weakened, indicating that part
phase through diffusion. As shown in Fig. 15, after 8 h of of the ­Mg54Ag17 phase was dissolved in the Mg matrix dur-
solution treatment, the peak corresponding to β is greatly ing the solution treatment. Combined with the distribution of
reduced, indicating that the β-Mg17Al12 phase has basically Ag elements in Fig. 16, there are still some ­Mg54Ag17 phase
dissolved into the α-Mg matrix. The holding time of solution particles distributed between the unsolidified block β phases.
treatment of AZ magnesium alloy obtained by traditional After 8 h of solution treatment, the existence of M ­ g54Ag17
casting process needs at least 20 h to fully solution [28]. phase was not detected, indicating that almost all ­Mg54Ag17
After 8 h of solution treatment for SLMed Mg–Al alloy, the phase was dissolved in α-Mg matrix.
β-Mg17Al12 phase has basically dissolved, because compared After T6 heat treatment, the microstructure of the
with as-cast Mg–Al alloy, the volume fraction of β-Mg17Al12 Mg–Al–Zn–2Ag alloy is shown in Fig. 16. EDS experi-
phase of laser additive manufacturing Mg–Al alloy is higher, ments on the stripped gray phases are shown in Fig. 17. At
but the distribution is more uniform and very thin, which the strip phases, the Mg element is significantly reduced
is beneficial to the diffusion of Al element. If continue to and the Al element content is increased. Combined with
extend the solution time, the grains may have a tendency to the analysis of literature data, this phase is precipitated
discontinuous β-Mg17Al12 phase during aging treatment
[28]. In Fig. 16a, it can be seen that there is a discontinu-
ous β-Mg17Al12 phase inside the grains, while the β phase
precipitated at the grain boundaries is granular. Figure 16b
is a magnification of Fig. 16a. The undissolved β-Mg17Al12
phase after solution treatment did not change its morphol-
ogy during aging treatment. As shown in Fig. 16c, there
are two forms of β-Mg17Al12 phase: unsolved β-Mg17Al12
phase and the discontinuous β-Mg17Al12 phase precipitated
during aging treatment. Figure 16d is an enlarged view
of the discontinuous β-Mg17Al12 phase inside the grain.
It can be seen that the discontinuous β phase is formed
by the interweaving and aggregation of the precipitated
short rod-shaped β-Mg 17Al 12, and the discontinuous β
phase has obvious directionality, and the direction of the
discontinuous β phase is different between different crystal
grains. During the aging heat treatment process, Al atoms
precipitate from the supersaturated α-Mg matrix and form
Fig. 15  XRD patterns of Mg–Al–Zn–2Ag alloy under different heat
­Mg17Al12 phase with Mg elements. At the grain boundary,
treatments

13
186 Page 12 of 16 C. Hu et al.

Fig. 16  Microstructure of Mg–Al–Zn–2Ag alloy after T6 heat treatment

Fig. 17  line scan analysis of EDS elements in Mg–Al–Zn–2Ag alloy after T6 treatment

the mismatch of atomic arrangement is larger, the energy The formation mechanism of the discontinuous β phase
is higher, and the β phase nucleates easily. As the β phase inside the crystal grain is the precipitation mechanism of
nucleates and grows at the grain boundary, the energy rises the cell zone. The β phase nucleates in the crystal grain
at the interface of β phase and matrix, and at the same near the grain boundary and grows into another grain. Dur-
time Al atoms are consumed to form an Al-poor zone at ing growing process, it pushes the grain boundary to move
the grain boundary, but the β phase cannot grow into the forward. Keep a certain direction and stop growing when
grain. Therefore, the granular β phase reaches equilibrium encountering the precipitated β phase [30].
when its diameter reaches 1–2 microns and stops growing.

13
Effects of Ag content and heat treatment on the microstructure and properties of SLMed AZ61 Mg–… Page 13 of 16 186

3.2.2 3.2.2 Mechanical properties of heat‑treated AZ61 Mg increases. When the content of Ag is 2%, the hardness of
alloys the sample without Ag is increased by 24.7%. This is due
to the aging hardening effect of Ag on Mg–Al alloy [23].
As shown in Fig. 18, the micro-hardness value of SLMed When the solution time is extended to 8 h, as the crystal
Mg–Al alloy after 6 h of solid solution is between 69.5 grains further grow, more β phase is dissolved in the matrix,
and 89.4 HV, which is significantly lower than the hard- and the average hardness value of each group of samples
ness before solid solution. When the Ag content is 1%, the is further reduced to between 68.1 and 76HV. As the Ag
hardness increased by 8.5%. This is due to the solid solu- content increases, the hardness increases slightly after solid
tion strengthening effect of the Ag element, but when the solution, showing the solid solution strengthening effect of
Ag content further increases, the hardness does not further Ag. Moreover, it can be seen from Fig. 18 that the hard-
increase, indicating that the effect of solid solution strength- ness value fluctuates less after 8 h of solid solution. This is
ening is limited. The hardness of Mg–Al alloy after aging because after 8 h of solid solution, the β phase is basically
is higher than that before heat treatment, and the hardness dissolved in the matrix, while after 6 h of solid solution, the
range is between 87.4 and 109.0 HV. And with the increase ununiformly distributed block β phase leads to slightly larger
in Ag content, the hardness of the sample after aging fluctuations of micro-hardness. After T6 treatment, age hard-
ening occurs, and the hardness of the two groups of aging is
similar. Although a longer time of solid solution leads to a
decrease in hardness, after 8 h of solid solution + aging treat-
ment, the Al element is basically solid-soluble in the α-Mg
matrix, which is conducive to the precipitation of ­Mg17Al12
phase during aging, and more formation of discontinuous
β-Mg17Al12 phase within grains improves the hardness of
Mg–Al alloy.
Nano-indentation experiments results are shown in
Fig. 19 and Table 6. The nano-hardness value decreases after
the solid solution experiment and increases after the aging
treatment, which is consistent with the micro-hardness value;
longer solution time leads to more β-Mg17Al12 phase dissolu-
tion, lower hardness, and more sufficient solid solution. It is
beneficial to the growth of discontinuous β-Mg17Al12 phase,
which increases the hardness after aging. After the solu-
tion test, the elastic modulus of the sample is reduced, and
the ability to resist deformation in the elastic deformation
Fig. 18  Micro-hardness of Mg–Al–Zn–Ag alloy samples after heat stage is weakened; the average residual depths of samples
treatment solutionized for 6 or 8 h are 226 and 230 nm, respectively,

Fig. 19  Nano-indentation results of Mg–Al–Zn–2Ag samples under different heat treatments a Load displacement curve; b Indentation depth
curve

13
186 Page 14 of 16 C. Hu et al.

Table 6  Elastic modulus and Heat treatment HIT (MPa) EIT (GPa) Penetration depth/Residual Rebound
indentation depth of Mg–Al– depth (nm) rate (%)
Zn–2Ag samples after different
heat treatments under 5 mN 2%Ag(as-deposited) 2224.3 48.62 264/192 72.7
load
6 h solution 1384.2 38.54 282/226 80.1
8 h solution 1285.3 38.83 296/230 77.7
6 h solution + aging 1921.5 46.72 275/168 61.1
8 h solution + aging 2886.8 53.09 225/118 52.4

which is greater than that of the unheated Mg–Al–Zn–2Ag


alloy sample. It shows that after solution treatment, the
plastic deformation increases and the alloys’ ability to resist
external force intrusion is weakened. As the solution time
increases, more β-Mg17Al12 phase is dissolved in the matrix,
and the plastic deformation increases further, indicating that
β-Mg17Al12 phase can resist deformation. After the aging
treatment, the plastic deformation was significantly reduced,
indicating that the re-precipitated β-Mg17Al12 phase can
effectively resist deformation. The residual depth and elas-
tic modulus of the sample treated with 8 h of solid solu-
tion + aging treatment are both the largest. This is because
Fig. 20  Residual depth curves of Mg–Al–Zn–2Ag alloy under differ-
the precipitated discontinuous β phase can effectively hinder ent heat treatment conditions
the movement of dislocations and improve the strength of
magnesium alloys [28], and 8 h of solid solution is more
conducive to the formation of discontinuous β-Mg 17Al12
phase. Because the heat treatment changes the microstruc-
ture of the SLMed Mg–Al alloys, the modulus of elasticity
changes significantly. After solution treatment, the rebound
rate increases; after aging, the rebound rate decreases. After
solution treatment, the phase that resists deformation of the
sample is mainly α-Mg matrix, and after aging, the sample
mainly depends on the discontinuous β phase to resist dis-
location movement and deformation, which may be because
the ability of α-Mg matrix to recover deformation is stronger
than of discontinuous β phase.
After different heat treatments, the Mg–Al–Zn–2Ag
alloy was subjected to a micron scratching experiment
Fig. 21  Curves of friction coefficient of Mg–Al–Zn–2Ag alloy under
under a load of 300 mN. The measured residual depth different heat treatment conditions
and friction coefficient curves are shown in Figs. 20 and
21. After 6 h and 8 h of solid solution experiment, the
average residual depth is 1869.64 nm and 1918.11 nm, average friction coefficients increased to 0.30 and 0.31,
respectively, which is larger than the residual depth of respectively, which were higher than the friction coeffi-
the unheated Mg–Al–Zn–2Ag alloy of 1114.96 nm. cient of the unheated sample 0.27; after aging treatment,
After aging treatment, the residual depth was reduced to the friction coefficients were reduced to 0.26, 0.19. The
807.80 nm and 730.25 nm, respectively. The same trend sample after solution treatment has less β-Mg17Al12 phase,
in the nano-indentation experiment shows that the dis- and the micron-scratching probe directly rubs against
continuous β-Mg17Al12 phase precipitated after the aging α-Mg matrix; the discontinuous β-Mg17Al12 phase precipi-
treatment can effectively resist microscopic deforma- tated after aging treatment can effectively reduce the fric-
tion and improve the hardness and resistance of Mg–Al tion coefficient and improve the wear resistance of Mg–Al
alloys. After 6 h and 8 h of solution test, the friction coef- alloy. Therefore, after 8 h of solution + aging treatment,
ficient of the samples increased to a certain extent. The more discontinuous β-Mg17Al12 phases are precipitated,
resulting in a lower friction coefficient.

13
Effects of Ag content and heat treatment on the microstructure and properties of SLMed AZ61 Mg–… Page 15 of 16 186

4 Conclusions

1. Adding a small amount of Ag element to the SLMed


Mg–Al alloy can inhibit the production of β-Mg17Al12
phase and further refine the grain size, and the formed
­M g 54Ag 17 strengthening phase is distributed at the
grain boundary. Addition of 2 wt% Ag can significantly
increase the micro-hardness of the SLMed Mg–Al alloy,
reduce deformation, and enhance corrosion resistance.
2. After 8 h of solution treatment, the β-Mg17Al12 phase
in the SLMed Mg–Al alloy is basically dissolved in the
matrix, and sufficient solid solution is more conducive
to the growth of discontinuous β-Mg17Al12 phase dur-
ing aging treatment. After solution treatment, the hard-
ness of Mg–Al alloy is decreased, and the ability to resist
external force intrusion is weakened, but the corrosion
resistance is significantly enhanced; after aging treatment,
Fig. 22  Polarization curves of Mg–Al–Zn–2Ag samples under differ- the micro-hardness and the ability to resist external force
ent heat treatments
intrusion are higher than that of unheated Mg–Al alloy,
but the corrosion resistance is slightly lower than that of
Table 7  Self-corrosion potential and corrosion current density of
unheated Mg–Al alloy.
Mg–Al–Zn–2Ag specimens under different heat treatments

Heat treatment Ecoor(vs. SCE)/V Jcoor/(A cm−2) Acknowledgements This work was supported by the Suzhou Science
–4 and Technology Bureau [Grant No. SYG201642], the Open Fund for
2%Ag(as-deposited) − 1.262 6.656 × 10
Jiangsu Key Laboratory of Advanced Manufacturing Technology
6 h solution − 1.175 5.424 × 10–4 [Grant No. HGAMTL-1701 and No. HGAMTL-1902], and the Jiangsu
8 h solution − 1.161 4.888 × 10–4 Province 333 Talent Project [Grant No. BRA2017098].
6 h solution + aging − 1.267 9.752 × 10–4
8 h solution + aging − 1.279 6.032 × 10–4 Compliance with ethical standards

Conflict of interest The authors declare that they have no conflict of


interest.
3.2.3 Corrosion resistance of heat‑treated Mg–Al–Zn–Ag
alloy

After the Mg–Al–Zn–2Ag sample is subjected to different References


heat treatments, the polarization curve and corrosion data
1. T.M. Pollock, Science 328, 986 (2010)
obtained through electrochemical experiments are shown 2. A.A. Luo, Int. Mater. Rev. 49, 13 (2004)
in Fig. 22 and Table 7. Compared with the unheated sam- 3. N. Kumar, M. Komarasamy, R.S. Mishra, J. Mater. Sci. 49, 4202–
ples, after 6 h and 8 h of solution treatment, the self-corro- 4214 (2014)
sion potential of the Mg–Al alloy samples increased from 4. C.E. Wen, M. Mabuchi, Y. Yamanda, K. Shimojima, Y. Chino, T.
Asahina, Scr. Mater. 45, 1147–1153 (2001)
− 1.262 to − 1.175 V and − 1.161 V, respectively, and the 5. J.P. Kruth, L. Froyen, J.V. Vaerenbergh, P. Mercelis, M. Rom-
corrosion resistance was enhanced. With the extension of bouts, B. Lauwers, J. Mater. Proc. Tech. 149(1), 616–622 (2004)
the solution time, the further reduction of the β-Mg17Al12 6. N. Gardan, A.J. Manuf. Syst., 37, 417–425 (2015)
phase increases the self-corrosion potential and improves 7. C.L. Mendis, K. Oh-ishi, Y. Kawamura, T. Honma, S. Kamado,
K. Hono, Acta Mater. 57, 749–760 (2009)
the corrosion resistance. After the aging treatment, the 8. C.C. Ng, M.M. Savalani, M.L. Lau, H.C. Man, Appl. Surf. Sci.
self-corrosion potential was reduced to − 1.267 V and 257, 7447–7454 (2011)
− 1.279 V, respectively, and the corrosion resistance was 9. S. Candan, M. Celik, E. Candan, J. Alloy. Compd. 672, 197–203
reduced, especially the corrosion resistance of the alloy (2016)
10. J. Feng, H. Sun, X. Li, H. Wang, W. Fang, J. Mater. Res. 31,
after 8 h of solid solution + aging was the worst, indicat- 3360–3371 (2016)
ing the precipitation of discontinuous β-Mg17Al12 phase 11. Y.M. Zhu, A.J. Morton, J.F. Nie, Scripta Mater. 58, 525 (2008)
It is not conducive to improving the corrosion resistance 12. J.S. Miao, W.H. Sun, A.D. Klarner, A.A. Luo, Scripta Mater. 154,
of magnesium aluminum alloy. 192–196 (2018)

13
186 Page 16 of 16 C. Hu et al.

13. K. Bryla, J. Horky, M. Krystian, L. Litynska-Dobrzynska, B. Min- 24. B.C. Zhang, H.L. Liao, C. Coddet, Mater. Des. 34, 753 (2012)
gler, Mater. Sci. Eng. C 109, 110543 (2020) 25. C. Vargel, M. Jacques, M.P. Schmidt, Corrosion of Aluminium
14. A.K. Dahle, Y.C. Lee, M.D. Nave, P.L. Schaffer, D.H. StJohn, J. (2004). https​://doi.org/10.1016/B978-0-08-04449​5-6.X5000​-9
Light Met. 1, 61–72 (2001) 26. L. Tan, T.R. Allen, Corros. Sci. 52(2), 548–554 (2010)
15. W.C. Zheng, S.-S. Li, B. Tang, D.B. Zeng, China Foundry 3(4), 27. D. Tie, F. Feyerabend, W.D. Müller et al., Eur. Mater. 25, 284–298
270–274 (2006) (2013)
16. V.E. Arakcheeva, A.V. Karpinskii, O.G. Koleanichenko, Krystal- 28. W. Zhou, T. Shen, N.N. Aung, Corros. Sci. 52(3), 1035–1041
lografiya 33, 907–908 (1988) (2010)
17. C. Dai, D.V. Malakhov, J. Alloys Compd. 619, 20–25 (2015) 29. X.U.E. Wenqi, C.H.E.N. Weihua, L.I.U. Juan et al., J. Plast. Eng.
18. Z.Z. Shi, W.Z. Zhang, Philos. Mag. Lett. 93, 473–480 (2013) 25(1), 15–21 (2018)
19. C.L. Mendis, C.J. Bettles, M.A. Gibson, S. Gorsse, C.R. Hutch- 30. T. Wei, H. Enhou, X.U. Yongdo et al. Acta metallurgica sinica,
inson, Philos. Mag. Lett. 86, 443–456 (2006) 0412-1961(2005)11-1199–08H
20. X.-F. Huang, W.-Z. Zhang, Mater. Sci. Eng. A 552, 211–221
(2012) Publisher’s Note Springer Nature remains neutral with regard to
21. E.A. Brandes, G. Brook, Smithells Metals Reference Book, 7th jurisdictional claims in published maps and institutional affiliations.
edn. (Butterworth-Heinemann, Oxford, 1992).
22. K.W. Wei, M. Gao, Z.M. Wang et al., Mater. Sci. Eng. A 611,
212–222 (2014)
23. J. Miao, W. Sun, A.D. Klarner et al., Scripta Mater. 154, 192–196
(2018)

13

You might also like