Sie sind auf Seite 1von 8

Materials Science and Engineering C 32 (2012) 2570–2577

Contents lists available at SciVerse ScienceDirect

Materials Science and Engineering C


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

Effects of Zn on microstructure, mechanical properties and corrosion


behavior of Mg–Zn alloys
Shuhua Cai a, Ting Lei a,⁎, Nianfeng Li a, b, Fangfang Feng a
a
State Key Laboratory of Powder Metallurgy, Central South University, Changsha 410083, China
b
Xiangya Hospital, Central South University, Changsha 410008, China

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

Article history: In this study, binary Mg–Zn alloys were fabricated with high-purity raw materials and by a clean melting pro-
Received 3 November 2011 cess. The effects of Zn on the microstructure, mechanical property and corrosion behavior of the as-cast Mg–
Received in revised form 3 July 2012 Zn alloys were studied using direct observations, tensile testing, immersion tests and electrochemical evalu-
Accepted 27 July 2012
ations. Results indicate that the microstructure of Mg–Zn alloys typically consists of primary α-Mg matrix
Available online 3 August 2012
and MgZn intermetallic phase mainly distributed along grain boundary. The improvement in mechanical per-
Keywords:
formances for Mg–Zn alloys with Zn content until 5% of weight is corresponding to fine grain strengthening,
Magnesium solid solution strengthening and second phase strengthening. Polarization test has shown the beneficial ef-
Biomaterials fect of Zn element on the formation of a protective film on the surface of alloys. Mg–5Zn alloy exhibits the
Electrochemical impedance spectra best anti-corrosion property. However, further increase of Zn content until 7% of weight deteriorates the cor-
Corrosion properties rosion rate which is driven by galvanic couple effect.
Mechanical property © 2012 Elsevier B.V. All rights reserved.

1. Introduction effects of metallic impurities, i.e. Fe and Ni [9]. Therefore, Zn-containing


Mg alloys have been paid more attention and developed as promising
Magnesium alloys have attracted great attention as an orthopedic candidates for biomedical applications.
biodegradable implant material due to their perfect biocompatibility Some recent studies discussed the mechanical property and corro-
and close mechanical properties to natural bone [1]. However, mag- sion behavior of binary Mg–Zn alloys. Zheng et al. [10] studied the
nesium alloys are extremely susceptible to corrosion, leading to losses binary Mg–1Zn alloy and reported its enhanced mechanical proper-
of strength and toughness, and thus limiting their practical applica- ties and corrosion resistance by addition of Zn as an alloying element.
tion as implant biomaterials [2]. Element alloying is one of the most A binary Mg–6Zn alloy revealed suitable tensile strength and elonga-
effective methods to improve the corrosion resistance and mechani- tion for implant application, as well as a reduced in vitro degradation
cal properties of magnesium. rate and good in vivo biocompatibility [11,12]. Also, the findings of
To date, most of the reported biomedical magnesium alloys con- Mg–3Zn alloy by different heat treatments showed that solution
tain aluminum and/or rare earth (RE) elements. Unfortunately, it is treatment enhanced corrosion resistance while aging treatment de-
found that the administration of Al and RE may induce latent toxic creased the corrosion resistance [13]. A study by Boehlert [14] on
and harmful effects on the human body [3–5]. Consequently, alloying Mg–Zn alloys containing 0–4.4 wt.% Zn suggests that Zn was a potent
elements must be chosen with careful consideration of the possible grain refiner and strengthener for Mg, where the optimal Zn content
toxic effects. Clearly, from a biological point of view, Al and RE are un- is 4 wt.%. All these aforementioned studies showed the enhanced cor-
suitable alloying elements, and thus it is necessary to develop a novel rosion resistance and mechanical property of magnesium by the incor-
biodegradable magnesium alloy without Al, RE or other harmful ele- poration of zinc in magnesium and great potential of binary Mg–Zn
ments for biomedical application. Song [6] explored in vitro corrosion alloys in biomedical applications. However, there is a lack of details on
rates of several magnesium alloys, pointing out that Ca, Mn and Zn the influence of volume fraction and existence format of secondary
could be appropriate candidates. MTT (3-(4,5-Dimethylthiazol- phases. Besides there were disagreements over the optimal Zn addition,
2-yl)-2,5-diphenyltetrazolium bromide) results also clearly indicated which was either low (1 wt.%) or high (6 wt.%).
that Mg, Zn and Ca did not have cytotoxicity [7]. It is also found that There is a high demand to design magnesium alloys with control-
zinc is one of the most abundant nutritionally essential elements in lable corrosion rates and suitable mechanical properties. Accordingly,
the human body, and has basic safety for biomedical applications it is necessary to study systematically the effects of zinc content on
[8]. Additionally, the addition of Zn can help to reduce the deleterious the microstructures, mechanical properties and degradation behavior
of binary Mg–Zn alloys for biomedical application. According to the
⁎ Corresponding author. Fax: +86 731 88710855. Mg–Zn binary phase diagram [15], the maximum solid solubility of
E-mail address: tlei@mail.csu.edu.cn (T. Lei). Zn in Mg is 6.2 wt.% (i.e. 2.5 at.%) at the eutectic temperature

0928-4931/$ – see front matter © 2012 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.msec.2012.07.042
S. Cai et al. / Materials Science and Engineering C 32 (2012) 2570–2577 2571

341 °C. For this reason, the alloys with different volume fractions and
existence formats of secondary phases are designed in this study by
the incorporation of Zn content lower than and close to its maximum
solid solubility. And thus three binary Mg–x wt.% Zn alloys (x= 1, 5, 7)
were prepared and investigated to evaluate their microstructure, me-
chanical property and corrosion behavior.

2. Experimental

Mg–Zn alloys, with Zn content of 1, 5 and 7 wt.%, were prepared


using high purity magnesium (≥99.99 wt.%) ingots and zinc
(≥99.99 wt.%) granules as raw materials. The melting is conducted
in an electronic resistance furnace at 750–800 °C protected by 99%
CO2 and 1% SF6 (volume fraction) mixed gas from oxidation. After
about 30 min holding and stirring, the melt is cast into a permanent
steel mould preheated to 200 °C to form an ingot. The chemical com-
Fig. 1. The photo of a cylindrical specimen for electrochemical measurements.
positions of the Mg–Zn alloys were analyzed by an inductively
coupled plasma atomic-emission spectrometry (ICP-AES) method,
as listed in Table 1. According to the nominal zinc contents, the where CR is the corrosion rate in mg cm −2 d −1, Δm is the weight loss
as-cast Mg–Zn alloy ingots were denoted as Mg–1Zn, Mg–5Zn and in mg, A is the original surface area exposed to the corrosive media in
Mg–7Zn, respectively. Pure Mg is used as control group. cm 2, and t is the immersion time in days. The weight loss was
Cylindrical specimens with a diameter of 13.5 mm and a height of converted to the average corrosion rate, Rw (mm/year) using [19]:
5 mm (Fig. 1) were machined by linear cutting and ground with SiC
emery papers up to 2000 grit, and successively polished with 1 μm Rw ¼ 2:1C R : ð2Þ
diamond paste, then ultrasonically cleaned in pure ethanol and
dried under an air pressure stream prior to experiments. Electrochemical measurements were performed on CHI‐660C
The Archimedes' principle was used to measure the density of electrochemical workstation with a three-electrode system compris-
Mg–Zn alloys and an average of three readings was taken for each ing the as-cleaned Mg–Zn alloy slice as working electrode by sealing
reported density. A dog-bone specimen with a gauge length of in a Teflon jacket with an exposed geometric area of 1 cm 2, a plati-
35 mm, a thickness of 2 mm and a width of 10 mm was machined num wire as auxiliary electrode, and a saturated calomel electrode
for tensile test [16]. Tensile strength was tested on an Instron 3369 (SCE) as reference electrode.
materials testing machine at a displacement rate of 2 mm min −1. Before the measurements, open circuit potential Eocp was tested
Three parallel specimens were taken for each group in the tensile until it was stabilized. Potentiodynamic polarization curves were
test. An extension meter with a gauge length of 10 mm was used to obtained in the potential range of Eocp ±200 mV at a scan rate of
measure the elongation. Cylindrical samples with a diameter of 2 mV s −1. The corrosion potential (Ecorr) and corrosion current density
10 mm and a height of 20 mm were used for compression test. Brinell (Icorr) were derived directly from the polarization curves by Tafel region
hardness measurements were performed using a Brinell hardness extrapolation. The corrosion current density, Icorr (mA/cm2), is related
tester (HB-RNu-1875) with a steel ball indenter of 2.5 mm in diame- to the corrosion rate (mm/year) using [19]:
ter under the load of 612.5 N and maintained for 30 s.
Electrochemical measurements and immersion tests were carried Ri ¼ 22:85Icorr : ð3Þ
out in simulated body fluid (SBF) [6]. The pH value of the solution was
adjusted to 7.4 at 37 ± 0.5 °C with 1.0 mol/L HCl and tris Electrochemical impedance spectroscopy (EIS) measurements
(hydroxymethyl) aminomethane (CH2OH)3CNH2 solution. Immer- were performed at Eocp with the scan frequency ranged from
sion tests were carried out in accordance with ASTMG31-72 [17] 100 kHz to 0.01 Hz, and with the perturbation amplitude of 5 mV.
(the ratio of surface area to solution volume was 1 cm 2:20 mL). The Specimens for optical microscopy were etched with a solution
pH value of the solution was recorded during the immersion tests consisting of 10 g picric acid, 175 mL ethanol, 25 mL acetic acid,
(PHS-3C pH meter, Lei-ci, Shanghai). The weight losses of the speci- and 25 mL distilled water. For the characterization of the sample
mens were obtained by immersion testing. After the immersion of morphology and composition, a field-emission scanning electron
5 days, the samples were removed from the solution and cleaned microscope Nova NanoSEM 230 equipped with an Energy disper-
with chromate acid (200 g/L CrO3 + 10 g/L AgNO3) to remove surface sive X-ray (EDX) analyzer was used. The phase composition of the
corrosion products without removing any amount of metallic Mg three Mg–Zn alloys was analyzed by X-ray diffractometry (XRD:
[18]. Then the samples were rinsed with distilled water, cleaned D/MAX-255) with the Cu–Kα1 radiation (wavelength λ =
ultrasonically in acetone, and dried in air. The dried specimens were 1.5406 Å). The tube voltage and the tube electric current of XRD
weighed and the corrosion rate (CR) can be calculated by Eq. (1) [17]: were 40 kW and 250 mA, respectively.

3. Results and discussion


C R ¼ Δm=At ð1Þ
3.1. Microstructures of the as-cast Mg–Zn alloys

Table 1
Fig. 2 displays the optical metallographic images of pure Mg and the
Chemical composition, wt.%, of pure Mg and the as-cast three Mg–Zn alloys.
as-cast Mg–Zn alloys with different Zn contents. It can be seen that the
Alloys name Zn Fe Cu Mn Mg microstructure of the as-cast Mg–Zn alloys typically consists of primary
Pure Mg 0.001 ≥99.9 α-Mg matrix and second phase mainly distributed along grain
Mg–1Zn 1.12 b0.0016 b0.002 b0.001 Balance boundary. It clearly shows that the grain size decreased with increasing
Mg–5Zn 5.10 Balance Zn content. The pure Mg material exhibits a grain size of 350 μm as
Mg–7Zn 7.05 Balance
shown in Fig. 2a, while the average grain sizes in Mg–1Zn, Mg–5Zn
2572 S. Cai et al. / Materials Science and Engineering C 32 (2012) 2570–2577

Fig. 2. Optical micrographs showing the microstructures of (a) pure Mg, (b) Mg–1Zn, (c) Mg–5Zn and (d) Mg–7Zn.

and Mg–7Zn are about 100 μm, 55 μm and 56 μm as observed in intermetallic, which explained the presence of MgZn phase in XRD
Fig. 2(b)–(d), respectively, indicating the addition of Zn element until analysis for Mg–5Zn and Mg–7Zn samples. Moreover, Zn has a rela-
5% of weight to Mg can significantly refine the grain size, but the refine- tively high solubility, and up to 1.6 wt.% Zn can fully dissolve in
ment efficiency is not significant with further addition of Zn over α-Mg matrix at room temperature [15]. Therefore, the alloying ele-
7 wt.%. When the Zn content increased from 5 to 7 wt.%, the second ment Zn tends to be incorporated into the α-Mg matrix when its ad-
phase formed a network structure of dentrite along the grain boundary, dition is below the solubility limit. Accordingly, Mg–1Zn was
as shown in Fig. 2d. The microstructure of pure Mg and Mg–Zn alloys essentially a single-phase alloy with Zn dissolved as a solute in the
was also imaged using backscattered electron (BSE) mode in SEM to re- α-Mg matrix as observed in Fig. 3b, and thus no MgZn phase was
veal compositional contrast, as shown in Fig. 3. Because of the large dif- detected at such a low Zn content.
ference in atomic number between Mg and Zn, such images clearly In this study, the microstructure of the as-cast Mg–Zn alloys clearly
reveal that pure Mg contains only α-Mg phase as shown in Fig. 3a, shows that the addition of Zn can significantly refine the grain size of
while Zn‐rich second phase can be detected in the microstructure of Mg matrix, which is consistent with the previous findings [20–23]. A
Mg–5Zn and Mg–7Zn alloys. As could be seen in Fig. 3c and d, the similar result was also reported in binary Mg–Y alloys, where the
white island-like regions are corresponding to the second phase distrib- grain size significantly decreased with increasing Y content [24]. As
uting at the grain boundary. The chemical composition of the grain area for the reason for the grain refinement, it was widely accepted that
was identified by EDS to be attributed to α-Mg matrix as shown in the segregation of zinc at the front of grain growth forms an intensive
Fig. 3e, while the second phase was identified to be composed of Mg constitutional undercooling in a diffusion layer ahead of the advancing
and Zn elements as shown in Fig. 3f, and the atomic ratio of Mg to Zn solid/liquid interface and then restricts the grain growth and promotes
was about 7:3, likely attributing to Mg7Zn3 phase. Moreover, the the nucleation of the primary Mg, and thus refines the grain size [25].
volume fraction of the second phase also increased with the increase The refinement efficiency of a solution element can be determined by
of Zn content from 5 to 7 wt.%. the calculation of a growth restriction factor (GRE) [26]. It has been
The X-ray diffraction (XRD) patterns of pure Mg and Mg–Zn alloys documented [26] that Zn had a higher GRE value (5.31) than Al (4.32)
are shown in Fig. 4. As could be seen in Fig. 4, only peaks correspond- and Y (1.70), meaning that Zn has more powerful growth restriction
ing to α-Mg matrix phase were found in the XRD pattern of pure Mg and better refinement efficiency.
sample, while MgZn peaks can be clearly identified in the as-cast Mg–
5Zn and Mg–7Zn samples. Furthermore, the diffraction intensity of 3.2. Mechanical properties
MgZn increased with increasing Zn content. According to the binary
alloy phase diagrams [15], Mg7Zn3 may conduct eutectic reaction at Table 2 summarizes the mechanical properties of the as-cast Mg–Zn
325 °C and decompose into α-Mg and MgZn intermetallic during alloys in comparison with the reported properties of natural bone [27].
cooling process. The high concentration of Mg atom in the second Compared to pure magnesium, the hardness and ultimate strength in
phase with a high atomic ratio of Mg to Zn (7:3) derived from EDS yield, tension and compression of Mg–Zn alloys increased with the
analysis is likely attributed to the existence of α-Mg in eutectic increase of Zn content until 5% of weight. On the contrary, Mg–7Zn
(MgZn + α-Mg). Accordingly, with respect to the results of XRD and alloy displays deteriorated mechanical properties. The elongation of
EDS, it is reasonable to conclude that the second phase is MgZn Mg–1Zn alloy exhibits the maximum value as high as 13.77%. When
S. Cai et al. / Materials Science and Engineering C 32 (2012) 2570–2577 2573

Fig. 3. BSE-mode SEM micrograph of the surface morphology of alloys (a) pure Mg, (b) Mg–1Zn, (c) Mg–5Zn and (d) Mg–7Zn; (e) EDS analysis corresponding to assigned A area; (f) EDS
analysis corresponding to assigned B area.

the Zn content was increased to 5 wt.%, the elongation was significantly MgZn phases will precipitate from Mg matrix along grain boundaries,
reduced to 8.5%. Further addition of Zn content until 7% of weight led to which promote the strength of Mg–Zn alloys by dispersion strength-
a serious drop in elongation to 6.0%, even less than the value of pure Mg. ening [14,30]. Thereby, fine grain strengthening, solid solution
It is noted that values in elongation of 8.4 and 18.8% were reported for strengthening and second phase strengthening contribute to the
Mg–4.4Zn and extruded Mg–6Zn alloy, respectively [11,14]. improvement in mechanical performances of Mg–Zn alloys with Zn
The increase in tensile strength and yield strength for Mg–Zn content until 5% of weight.
alloys with Zn content until 5% of weight can be explained, on the Moreover, many studies [30,31] have pointed out that precipitated
one hand, by Hall–Patch relationship [28], which is supported by a second phase can dramatically improve the strength while decrease
number of findings [20–23]. On the other hand, according to the the plastic of alloys. On the one hand, the second phase may hinder
Mg–Zn binary phase diagram [15], the maximum solubility of Zn in the dislocation reduction and increase the dislocation density. On
magnesium is 1.6 wt.% at room temperature in the equilibrium the other hand, the second phase dispersed at the grain boundary
state, and thus Zn element mainly dissolves into primary Mg to could be new crack source, which expands easily and eventually re-
some extent, generating solid-solution strengthening [14,27,29]. In sults in brittle failure. As could be seen in Fig. 2d, plenty of second
addition, when the addition of Zn content is 5 wt.%, a number of phases in Mg–7Zn alloy formed a network structure with dendritic
2574 S. Cai et al. / Materials Science and Engineering C 32 (2012) 2570–2577

11.0
Mg
∇ MgZn 10.5

10.0
Mg-7Zn ∇ ∇
Intensity(a.u.)

9.5

pH value
Mg-5Zn ∇∇ 9.0

8.5
Pure Mg
Mg-1Zn
Mg-1Zn
8.0 Mg-5Zn
Mg-7Zn
Pure Mg 7.5

20 30 40 50 60 70 80 7.0
0 10 20 30 40 50 60 70
Angle (2θ)
immersion time(h)
Fig. 4. X-ray diffraction patterns of pure Mg and as-cast Mg–Zn alloys.
Fig. 5. The pH value of SBF as a function of immersion time for pure Mg and Mg–Zn
alloys.
segregation along grain boundaries, resulting in residual defects, and
thus sharply aggravated the strength and elongation of alloy. Overall,
Zn is a promising alloying element and more importantly, the addi- numbers of hydrogen bubbles are evidently observed arising from
tion of Zn content has to be controlled to promote the mechanical the surface of pure Mg specimens, indicating a fast rate of hydrogen
properties of Mg alloys to make them more suitable for human im- evolution due to the reaction of Mg matrix with the corrosive electro-
plant materials. lyte, while the numbers of hydrogen bubbles arising from the surfaces
of the three Mg–Zn alloys are comparatively fewer. After a few hours,
the pure Mg samples were covered with white corrosion products. As
3.3. Immersion test
the immersion time increased, corrosion products are observed on all
specimens and are increased with time. It is noted that detached cor-
The pH variation of SBF as a function of immersion time is shown in
rosion products are found on the bottom of the beaker containing
Fig. 5. It can be seen that the pH values of the solution corresponding to
pure Mg specimens at the end of immersion testing.
pure Mg specimens increased rapidly from 7.4 to 10.3 in the initial 10 h
The average weight changes of pure Mg and Mg–Zn alloys after
of immersion. Afterwards, the pH value of SBF increased slowly with
5 days of immersion in SBF were depicted in Fig. 6. The weight loss
immersion time and became stabilized at 10.2. In contrast, the pH values
is an indication that there was corrosion attack on all specimens.
of the solution corresponding to Mg–1Zn, Mg–5Zn and Mg–7Zn speci-
The maximum weight loss of 0.69 mg cm −2 h −1 was observed for
mens increased much slowly and reached 9.16, 9.04 and 9.17 in the ini-
pure Mg specimens, while the order of mass loss rate from high to
tial 10 h of immersion, and finally stabilized at 9.57, 9.46 and 9.67 at the
low is 0.063, 0.04 and 0.025 mg cm −2 h −1 for Mg–7Zn, Mg–1Zn
end of the immersion testing, respectively.
and Mg–5Zn, respectively. The correlated average corrosion rate, Rw
It is well known that corrosion behaviors of Mg and its alloy are
(mm/year) is listed in Table 3. The immersion result clearly indicates
correlated to their microstructures and corrosion starts on the α-Mg
that the corrosion resistance of Mg–Zn alloys increases with the in-
matrix phase. Mg is rather active in aqueous medium and dissolves
crease of Zn addition until 5% of weight. Excessive Zn content in
according to the following reaction [32]:
Mg–Zn alloy until 7% of weight leads to a serious drop in corrosion
þ 2þ − resistance.
Mg þ H þ H2 O→Mg þ OH þ H2 : ð4Þ
In order to get a better insight on the corroded surface appearance of
Consequently, in the early stage of immersion, the dissolution of all specimens after 5 days immersion testing, the corrosion products
magnesium consumes H +, but releases OH −, leading to the increase were removed by chromic acid. Fig. 7 showed the SEM surface
of pH value of SBF [16], and thus the change of pH value with immer-
sion time could be used to evaluate the corrosion behavior of magne-
0.7
sium. It is reasonable to deduce from the pH variation trend that all Pure Mg
average weight loss (mg/cm2/h)

three Mg–Zn alloys exhibit a relatively slower degradation rate as


0.6
compared to pure Mg material.
The corrosion behavior of the three Mg–Zn alloys was also evalu- 0.5
ated by immersion test in SBF. In the early stage of immersion, large
0.4

Table 2
0.3
Mechanical properties of the as-cast three Mg–Zn alloys.

Material Modulus Yield Tensile Elongation Compression Hardness 0.2


(GPa) strength strength (%) strength (HB)
(MPa) (MPa) (MPa) 0.1 Mg-1Zn Mg-7Zn
Mg-5Zn
Natural 5–23 – 35–283 1.07–2.10 164–240 –
bone 0.0
[28] -1 0 1 2 3 4 5 6 7 8 9
Pure Mg 1.86 29.88 100.47 7.43 183.09 37.10 Zn content(wt.%)
Mg–1Zn 24.23 60.62 187.73 13.77 329.60 47.33
Mg–5Zn 36.47 75.60 194.59 8.50 334.12 53.80
Fig. 6. Average corrosion rates determined from weight loss testing for pure Mg and
Mg–7Zn 39.60 67.28 135.53 6.00 353.11 56.26
Mg–Zn alloys.
S. Cai et al. / Materials Science and Engineering C 32 (2012) 2570–2577 2575

Table 3 3.4. Electrochemical corrosion measurements


Values measured from weight loss and the polarization curves for pure Mg and Mg–Zn
alloys in SBF.
The corrosion resistance of three Mg–Zn alloys was further deter-
Material Ecorr (VSCE) Icorr (μA/cm2) Ri (mm/y) Eb (VSCE) Rw (mm/y) mined in SBF using potentiodynamic polarization test as shown in
Pure Mg −1.581 680.1 15.30 – 34.78 Fig. 8. In general, the cathodic polarization curve is attributed to hy-
Mg–1Zn −1.527 23.4 0.53 −1.09 2.01 drogen evolution reaction due to the reduction of water, while the
Mg–5Zn −1.477 11.72 0.26 −0.96 1.26
anodic polarization curve is associated with the dissolution of Mg,
Mg–7Zn −1.543 51.79 1.17 −1.25 3.18
leading to the formation of Mg 2+ [32]. The corrosion potential
(Ecorr) and corrosion current density (Icorr) were derived directly
from the polarization curves by Tafel region extrapolation and were
morphologies of the as-cleaned pure Mg and Mg–Zn alloy specimens. listed in Table 3. The corrosion potential of the as-cast Mg–1Zn and
As can be seen in Fig. 7a, the surface of pure Mg collapsed severely Mg–5Zn alloy is − 1.53 and − 1.47 V, meanwhile decrease in the cor-
and showed lamellar microstructures, indicating a substantial corrosion rosion current density (Icorr) is observed, whereas Ecorr of the as-cast
rate during the immersion testing. Fig. 7(b)–(d) presented the surface Mg–7Zn alloy is − 1.54 V. Apparently, the addition of the Zn element
appearances of three Mg–Zn alloys, which were characterized by a until 5% of weight shifts the corrosion potential toward noble direc-
number of deep pits (as indicated by arrows) superimposed on superfi- tion. It is clearly deduced from Table 3 that all three Mg–Zn alloys
cial corrosion over the whole specimen surfaces, implying localized cor- exhibit superior corrosion resistance to pure Mg material. The in-
rosion attacks occurred during the immersion process. It is worthy to crease in degradation rate of the alloys is in the following order:
note the presence of pits is detrimental to the overall corrosion resis- Mg–5Zn, Mg–1Zn and Mg–7Zn. It has been reported that zinc is able
tance and the destruction of Mg alloys will proceed by means of pitting to elevate the corrosion potential of magnesium alloys, and thus re-
corrosion from one or more of them [33]. Mg–5Zn alloy exhibited duce the corrosion rate [34], which was in agreement with this study.
slighter pitting corrosion among all three alloys, whereas severe pitting Furthermore, a passivation stage was found in the anodic polariza-
corrosions accompanied by collapses over the whole surface of Mg–7Zn tion curve of all three Mg–Zn alloys. The breakdown potential (Eb) is
alloy were observed, indicating the slowest corrosion rate of Mg–5Zn usually indicated by a sudden drop on the polarization curve, which is
alloy and substantial heterogeneous corrosion of Mg–7Zn alloy. an indication of the tendency for localized corrosion. A more positive

Fig. 7. SEM microstructure of corroded appearance after 5 days immersion and after corrosion product removal for (a) pure Mg, (b) Mg–1Zn, (c) Mg–5Zn and (d) Mg–7Zn.
2576 S. Cai et al. / Materials Science and Engineering C 32 (2012) 2570–2577

-1.5 The significant difference in microstructure between them is the vol-


-2.0 ume fraction of the second phase. In general, second phase containing
log Current Density ( A / cm2)

-2.5 binary alloy of Mg and Zn has nobler potential than Mg matrix [37].
a
-3.0 Eb As a result, second phase may act as a cathode and Mg matrix as an
-3.5 anode at the interface between them, resulting in galvanic corrosion.
-4.0
Eb Correspondingly, the discrepancy in corrosion resistance for Mg–5Zn
d
b Eb
and Mg–7Zn is likely attributed to the role of second phases in corro-
-4.5
c sion process. The influences of second phases on the corrosion rate of
-5.0
alloys have been addressed recently [37,38]. In order to confirm the
-5.5 (a) pure Mg
effect of second phases, the cathodic polarization curve of Mg–7Zn
(b) Mg-1Zn
-6.0 was compared with that of Mg–5Zn. The cathodic current density of
(c) Mg-5Zn
-6.5 (d) Mg-7Zn both curves gradually increased with the enhancement of cathodic
-7.0 potential. The curve of the cathodic region to some extent represents
-7.5 polarization behavior of the non-corroded surface of a specimen and
-1.8 -1.6 -1.4 -1.2 -1.0 -0.8 -0.6 the reaction of hydrogen evolution [39]. At the same cathodic poten-
tial (marked by the broken line in Fig. 8), the corrosion current den-
Potential (V vs. SCE) sity of Mg–7Zn was nearly two orders of magnitude higher than
that of Mg–5Zn, indicating the fast cathodic hydrogen evolution rate
Fig. 8. Potentiodynamic polarization curves of pure Mg and Mg–Zn alloys.
on Mg–7Zn alloy. A similar result was also reported in Mg–Zn–Y–Zr
alloy with 5.8 wt.% Zn content [37]. The accelerated hydrogen evolu-
Eb means a less likely localized corrosion [35]. A breakdown potential tion rate can be attributed to the presence of high volume percent
at which pitting corrosion initiated was detected at − 1.09, − 0.95 second phases in Mg–7Zn alloy, which is in the form of a continuous
and − 1.25 V for Mg–1Zn, Mg–5Zn and Mg–7Zn alloy, respectively, network structure as shown in Fig. 2d. With the increase of Zn con-
as shown in Table 3. Therefore, the localized corrosion can more eas- tent over 5 wt.%, the quantity of MgZn intermetallic phase is in-
ily occur on the Mg–7Zn alloy. In contrast with the three alloys, pure creased rapidly in the matrix to form a continuous network
magnesium was active over all the potential range. The passivation structure, leading to the formation of more anode–cathode sites. Thus,
stage suggests that the Zn element could help to form a protective more galvanic corrosions could happen on these sites, resulting in fast
film on the surface of Mg–Zn alloys during corrosion process [36]. It corrosion rate. Accordingly, it can be inferred that corrosion of Mg–
is obvious that Mg–5Zn alloy shows the lowest anodic current and 7Zn alloy is driven by galvanic couple effect, resulting in the fast disso-
noblest corrosion potential followed by a simultaneous increase of lution of α-Mg matrix and thus the decline of corrosion resistance.
the breakdown potential, indicative of the best corrosion resistance Electrochemical impedance spectra (EIS) measurements for the
among the three alloys. This conclusion is in good agreement with three Mg–Zn alloys were carried out at open circuit potential (Eocp)
the immersion results. as illustrated in Fig. 9a. Apparently, the impedance diagram of pure
As confirmed in the microstructure characterization, both Mg–5Zn Mg is characteristic of one well-defined capacitive loop. In contrast,
and Mg–7Zn alloys are mainly composed of primary Mg matrix phase the EIS spectra of the three Mg–Zn alloys are characterized by two
with similar grain size and second phase distributed along grain loops: a capacitive loop in the high frequency region (HF) and a ca-
boundary with the same chemical composition of MgZn intermetallic. pacitive loop in the medium frequency region (MF), as labeled. In
general, the high frequency capacitive loop is attributed to the relax-
ation process of electrochemical reaction impedance corresponding
a to the dissolution of Mg and electric double layer capacitance (Cdl)
Rt (Ω cm2) pure Mg
1000 pure Mg Mg-7Zn Mg-1Zn Mg-5Zn at the interface between the metal surface and the corrosive medium
Mg-1Zn
914 1317 2085 2213 Mg-5Zn [40]. The capacitive loop observed at the middle frequency region is
800 HF
Mg-7Zn related to the presence of surface film influencing the corrosion pro-
cess [24]. Thereby, the electrode reaction process correlated to high
Z'' (ohm cm2)

frequency capacitive loop can be described by a parallel circuit of Cdl


600
MF and charge transfer resistance Rt [40]. The EIS plots for pure Mg and
Mg–Zn alloys were different in shape and diameter of the loops, in-
400 dicative of different corrosion mechanisms and corrosion rates. The
Nyquist plots of Mg–Zn alloys can be interpreted using the equivalent
200 circuit shown in Fig. 9b. Since the Nyquist plots of Mg–Zn alloys ex-
hibit depressed semicircles, a constant phase element, CPE1 is used
0 instead of Cdl in the proposed model. CPE2 and a resistance Rf are
also introduced to account for the capacitance and resistance of the
0 400 800 1200 1600 2000 2400 2800 corrosion product layer formed on the surface of Mg–Zn alloys. The
Rf displays correlation with the passivation stage observed on anodic
Z' (ohm cm2)
polarization curve due to protection layer formation. The diameter of
b the high frequency semicircle gives the charge-transfer resistance
(Rt) at the electrode/electrolyte interface. From Rt value, the
exchange-current density (j0) could be calculated using the following
expression [41]:

j0 ¼ RT=nFRt ð5Þ

where n is the number of transferred charges, F is Faraday constant.


Fig. 9. (a) Nyquist plots for pure Mg and Mg–Zn alloys after soaking in SBF for 1 h and Apparently, j0 is in inverse proportion to Rt, in other words, the higher
(b) equivalent circuit used for modeling experimental EIS data of Mg–Zn alloys. the Rt is, the lower would be the corrosion rate [41]. Consequently,
S. Cai et al. / Materials Science and Engineering C 32 (2012) 2570–2577 2577

charge transfer resistance could be used to evaluate the corrosion [4] C.H. Ku, D.P. Pioletti, M. Browne, P.J. Gregson, Biomaterials 23 (6) (2002)
1447–1454.
property of the alloys. This is because the larger the Rt is, the more dif- [5] N. Yumiko, T. Yukari, T. Yasuhide, S. Tadashi, I. Yoshio, Fundam. Appl. Toxicol. 37
ficult is the transfer of charges between solution and the sample sur- (1997) 106–116.
face, and thus the corrosion rate is low. It can be deduced from the [6] G. Song, Corros. Sci. 49 (4) (2007) 1696–1701.
[7] L. Yang, E.L. Zhang, Mater. Sci. Eng. C 29 (2009) 1691.
Nyquist plots that Rt ranked from low to high as following: pure Mg [8] H. Tapiero, K.D. Tew, Biomed. Pharmacother. 57 (9) (2003) 399–411.
(914 Ω cm 2), Mg–7Zn (1317 Ω cm 2), Mg–1Zn (2085 Ω cm 2) and [9] G.Q. Li, G.H. Wu, Y. Fan, W.J. Ding, Foundry Technol. 27 (2006) 81–82.
Mg–5Zn (2213 Ω cm 2). Meanwhile, the order of film resistance Rf [10] X.N. Gu, Y.F. Zheng, Y. Cheng, S.P. Zhong, T.F. Xi, Biomaterials 30 (2009) 484–498.
[11] S.X. Zhang, X.N. Zhang, C.L. Zhao, J.N. Li, Y. Song, C.Y. Xie, et al., Acta Biomater. 6
from high to low is 736, 623 and 549 Ω cm2 for Mg–5Zn, Mg–1Zn and (2010) 626–640.
Mg–7Zn alloy, respectively, which further indicates that Mg–5Zn alloy [12] S.X. Zhang, J.N. Li, Y. Song, C.L. Zhao, X.N. Zhang, C.Y. Xie, et al., Mater. Sci. Eng. C
has superior corrosion property. As a result, Mg–5Zn alloy exhibits the 29 (2009) 1907–1912.
[13] X.B. Liu, D.Y. Shan, Y.W. Song, E.H. Han, Trans. Nonferrous Met. Soc. China 20
best corrosion resistance, and the corrosion tendency displays well cor-
(2010) 1345–1350.
relation with the polarization measurement and immersion tests. [14] C.J. Boehlert, K. Knittel, Mater. Sci. Eng., A 417 (2006) 315–321.
[15] T.B. Massalski, Binary Alloy Phase Diagrams, American Society for Metals, Ohio,
4. Conclusions 1986.
[16] E.L. Zhang, L. Yang, J.W. Xu, H.Y. Chen, Acta Biomater. 6 (2010) 1756–1762.
[17] In: Annual Book of ASTM Standards[S], American Society for Testing and
Microstructure observation showed that the addition of Zn ele- Materials, Philadelphia, PA, 2004.
ment significantly refined the grain size of as-cast Mg–Zn alloys and [18] M.C. Zhao, P. Schmutz, S. Brunner, M. Liu, G.L. Song, A. Atrens, Corros. Sci. 51
(2009) 1277–1292.
increased the volume percent intermetallic in the microstructure. [19] Z.M. Shi, A. Atrens, Corros. Sci. 53 (2011) 226–246.
The tensile properties and corrosion resistance of the Mg–Zn alloys [20] F. Rosalbino, S. De Negri, A. Saccone, E. Angelini, S. Delfino, J. Mater. Sci. Mater.
strongly depended on the volume fraction and existence format of Med. 21 (2010) 1091–1098.
[21] E.L. Zhang, L. Yang, Mater. Sci. Eng., A 497 (2008) 111–118.
secondary phases. The corrosion morphologies and immersion testing [22] E.L. Zhang, W.W. He, H. Du, K. Yang, Mater. Sci. Eng., A 488 (2008) 102–111.
as well as electrochemical measurements proved that the corrosion [23] H. Du, Z.J. Wei, X.W. Liu, E.L. Zhang, Mater. Chem. Phys. 125 (2011) 568–575.
resistance increased with increasing Zn content in the range [24] M. Liu, P. Schmutz, P.J. Uggowitzer, G.L. Song, A. Atrens, Corros. Sci. 52 (2010)
3687–3701.
1–5 wt.% despite the presence of the potentially detrimental MgZn [25] E.L. Zhang, D.S. Yin, L.P. Xu, L. Yang, K. Yang, Mater. Sci. Eng. C 29 (2009) 987–993.
intermetallic. Excessive addition of Zn over 7 wt.% resulted in a net- [26] Y.C. Lee, A.K. Dahle, S.H. StJohn, Metall. Mater. Trans. A 31 (11) (2000)
work structure of MgZn intermetallic as a cathode, causing micro- 2895–2906.
[27] F. Witte, N. Hort, C. Vogt, S. Cohen, K.U. Kainer, R. Willumeit, F. Feyerabend, Curr.
galvanic corrosion acceleration. It was suggested that the addition of
Opin. Solid State Mater. Sci. 12 (2008) 63–72.
Zn content has to be precisely controlled to promote both the me- [28] E.O. Hall, Yield Point Phenomena in Metals and Alloys, Macmillan and Co. Ltd.,
chanical properties and corrosion resistance of Mg alloys for biomed- London, 1970.
ical application. [29] H. Somekawa, Y. Osawa, T. Mukai, Scr. Mater. 55 (2006) 593–596.
[30] D.S. Yin, E.L. Zhang, S.Y. Zeng, Trans. Nonferrous Met. Soc. China 18 (2008)
763–768.
Acknowledgment [31] D.F. Zhang, X.B. Zhao, G.L. Shi, Mater. Sci. Forum 610–613 (2009) 880–883.
[32] Y. Wang, M. Wei, J.C. Gao, J.Z. Hu, Y. Zhang, Mater. Lett. 62 (2008) 2181–2184.
[33] R. Hahn, J.G. Brunner, J. Kunze, P. Schmuki, S. Virtanen, Electrochem. Commun. 10
This work was supported by National Natural Science Foundation (2008) 288–292.
of China (grant no. 51021063), National Science Fund for Distin- [34] Z. Shi, G. Song, A. Atrens, Surf. Coat. Technol. 201 (2006) 492–503.
guished Young Scholars (grant no. 50825102) and Open Project of [35] J.W. Chang, P.H. Fu, X.W. Guo, L.M. Peng, W.J. Ding, Corros. Sci. 49 (6) (2007)
2612.
State Key Laboratory for Powder Metallurgy of CSU. [36] G.Y. Li, J.S. Lian, L.Y. Niu, Z.H. Jiang, Q. Jiang, Surf. Coat. Technol. 201 (2006) 1814.
[37] Y.W. Song, D.Y. Shan, R. Chen, E.H. Han, Corros. Sci. 52 (2010) 1830–1837.
References [38] W.W. He, E.L. Zhang, K. Yang, Mater. Sci. Eng. C 30 (2010) 167–174.
[39] G.L. Song, A. Atrens, X.L. Wu, B. Zhang, Corros. Sci. 40 (1998) 1769–1791.
[1] M.P. Staiger, A.M. Pietaka, J. Huadmaia, G. Dias, Biomaterials 27 (2006) 1728. [40] G. Song, A. Atrens, D. St John, X. Wu, J. Nairn, Corros. Sci. 39 (1997) 1981–2004.
[2] G. Song, A. Atrens, Adv. Eng. Mater. 5 (2003) 837. [41] R. Udhayan, P.B. Devendra, J. Power Sources 63 (1996) 103–107.
[3] S.S.A. El-Rahman, Pharmacol. Res. 47 (3) (2003) 189–194.

Das könnte Ihnen auch gefallen