You are on page 1of 15

Acta Biomaterialia 6 (2010) 626–640

Contents lists available at ScienceDirect

Acta Biomaterialia
journal homepage: www.elsevier.com/locate/actabiomat

Research on an Mg–Zn alloy as a degradable biomaterial


Shaoxiang Zhang a, Xiaonong Zhang a,c,*, Changli Zhao a, Jianan Li a, Yang Song a, Chaoying Xie a,
Hairong Tao b, Yan Zhang b, Yaohua He b, Yao Jiang b, Yujun Bian c
a
State Key Laboratory of Metal Matrix Composites, School of Materials Science and Engineering, Shanghai Jiao Tong University, Shanghai 200240, People’s Republic of China
b
Orthopaedic Department of the 6th People’s Hospital, Shanghai Jiao Tong University, Shanghai 200233, People’s Republic of China
c
Shanghai Origin Material and Medical Technology Co. Ltd., Shanghai 200240, People’s Republic of 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 a binary Mg–Zn magnesium alloy was researched as a degradable biomedical material. An
Received 17 December 2008 Mg–Zn alloy fabricated with high-purity raw materials and using a clean melting process had very low
Received in revised form 1 June 2009 levels of impurities. After solid solution treatment and hot working the grain size of the Mg–Zn alloy
Accepted 11 June 2009
was finer and a uniform single phase was gained. The mechanical properties of this Mg–Zn alloy were
Available online 21 June 2009
suitable for implant applications, i.e. the tensile strength and elongation achieved were 279.5 MPa
and 18.8%, respectively.
Keywords:
The results of in vitro degradation experiments including electrochemical measurements and immer-
Mg–Zn alloy
Degradation
sion tests revealed that the zinc could elevate the corrosion potential of Mg in simulated body fluid (SBF)
Mechanical properties and reduce the degradation rate. The corrosion products on the surface of Mg–Zn were hydroxyapatite
Cytotoxicity (HA) and other Mg/Ca phosphates in SBF. In addition, the influence caused by in vitro degradation on
In vivo biocompatibility mechanical properties was studied, and the results showed that the bending strength of Mg–Zn alloy
dropped sharply in the earlier stage of degradation, while smoothly during the later period.
The in vitro cytotoxicity of Mg–Zn was examined. The result 0–1 grade revealed that the Mg–Zn alloy
was harmless to L-929 cells. For in vivo experiments, Mg–Zn rods were implanted into the femoral shaft
of rabbits. The radiographs illustrated that the magnesium alloy could be gradually absorbed in vivo at
about 2.32 mm/yr degradation rate obtained by weight loss method. Hematoxylin and eosin (HE) stained
section around Mg–Zn rods suggested that there were newly formed bone surrounding the implant.
HE stained tissue (containing heart, liver, kidney and spleen tissues) and the biochemical measure-
ments, including serum magnesium, serum creatinine (CREA), blood urea nitrogen (BUN), glutamic-pyru-
vic transaminase (GPT) and creatine kinase (CK) proved that the in vivo degradation of Mg–Zn did not
harm the important organs. Moreover, no adverse effects of hydrogen generated by degradation had been
observed and also no negative effects caused by the release of zinc were detected. These results suggested
that the novel Mg–Zn binary alloy had good biocompatibility in vivo.
Ó 2009 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.

1. Introduction 4D-lactide, the magnesium alloys AZ31 and AZ91 enhanced the
osteogenesis response and increase newly formed bone [4]. Heub-
Due to their very low corrosion potential, magnesium and its lein et al. reported that the alloy AE21 gradually degraded in pig ar-
alloys are susceptible to dissolution in aqueous solutions, particu- tery and might be a promising cardiovascular implant material [3].
larly in those containing chloride ion electrolytes [1,2]. Making use A more encouraging indication is that a biodegradable magnesium
of the corrodible properties of magnesium alloys, in recent years stent has been used in clinical experiments [18].
biomaterials engineers have become interested in developing mag- However, it should be noted that most of the reported biomed-
nesium-based biodegradable medical devices [3–18]. ical magnesium alloys contain aluminum and/or rare earth (RE)
As shown in Table 1, various magnesium alloys have been re- elements. It is well known that Al is harmful to neurons [19] and
searched as biodegradable materials and some of them have osteoblasts [20] and also associated with dementia and Alzhei-
shown biocompatibility. For example, compared with poly-96L/ mer’s disease [19]. The administration of RE (Pr, Ce, Y, etc.) could
lead to hepatotoxicity [21]. Excessive yttrium ions (Y3+) have been
* Corresponding author. Address: State Key Laboratory of Metal Matrix Compos- shown to change the expression of some rat genes and to have ad-
ites, School of Materials Science and Engineering, Shanghai Jiao Tong University, No. verse effects on DNA transcription factors [22]. Consequently, Al
800 DongChun Road, Shanghai 200240, People’s Republic of China. Tel./fax: +86 21 and RE are unsuitable alloying elements for biomedical magnesium
3420 2759.
materials, particularly when they are above certain levels. This has
E-mail address: xnzhang@sjtu.edu.cn (X. Zhang).

1742-7061/$ - see front matter Ó 2009 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.
doi:10.1016/j.actbio.2009.06.028
S. Zhang et al. / Acta Biomaterialia 6 (2010) 626–640 627

Table 1
The nominal chemical compositions of Mg alloys researched as degradable biomaterials.

Alloy type Aluminum (wt.%) Rare earth element Zinc (wt.%) Other References
AE21 2 1 wt.% (Ce, Pr, Nd) [3]
AZ31 3 1 [4]
AZ91 9 1 [4,5,8,10]
WE43 3 wt.% (Nd, Ce, Dy), 4 wt.% (Y) [4]
LAE442 4 2 wt.% (Ce, La, Nd, Pr) 4 wt.% Li [4,5]
AM60B 6 0.07 0.33 wt.% Mn [14]
Mg–Ca <2 wt.% Ca [6,7,17]
Mg–Zn–Y 0.36–1.54 wt.% Y 2 [11]
Mg–Mn–Zn 1 1.2 wt.% Mn [16]

led to a demand for the development of a novel biodegradable 2.2. Composition and microstructure characterization
magnesium alloy without Al, RE or other harmful elements.
With the purpose of searching for suitable alloying elements for The chemical compositions of the Mg–6Zn alloy was deter-
biomedical magnesium alloys, Song [12] explored in vitro corro- mined by inductively coupled plasma-atomic emission spectrome-
sion rates of several magnesium alloys, pointing out that Ca, Mn try (ICP-AES) (Iris Advantage 1000). The microstructures of the as
and Zn could be appropriate candidates. Further investigations cast, heat-treated and extruded samples were characterized by
demonstrated that Mg–Ca [6] and Mg–Mn–Zn [16] alloys gradually optical microscopy. X-ray diffraction analysis (XRD) (Rigaku
degraded within bone and had good biocompatibility both in vitro D/MAX255) was used to examine the phase of the as cast, heat-
and in vivo. Moreover, the degradation process did not raise the treated and extruded Mg–6Zn.
serum Mg2+ level and no kidney disorders were observed.
Zinc is one of the most abundant nutritionally essential ele- 2.3. Mechanical properties
ments in the human body [23], and has basic safety for biomedical
applications. Furthermore, zinc can improve the corrosion resis- Tension and compression tests were carried out with a MTS
tance and mechanical properties of magnesium alloys. For exam- 810.22 universal testing machine, according to ASTM E8-04 and
ple, the corrosion rate of magnesium could be reduced by ASTM E9-89a (2000) [28,29]. The tensile samples had a gauge length
increasing the mass fraction of zinc in magnesium [24]. Moreover, of 35 mm and the compressive samples had diameter of 6 mm and a
zinc can effectively strengthen magnesium [12,25] through a solid height of 9 mm. A crosshead speed of 0.5 mm min1 was used.
solution hardening mechanism [25].
According to the Mg–Zn binary phase diagram [26] the maximum 2.4. In vitro degradation tests
solubility of Zn in Mg is 6.2 wt.% (i.e. 2.5 at.%) at 325 °C. In this paper a
patented Mg–6 wt.% Zn alloy (Mg–6Zn) [27] was developed with the In order to evaluate the in vitro degradation properties, electro-
aim of making a novel magnesium alloy with good biocompatibility, chemical measurements and immersion tests were performed in a
moderate degradation rate and good mechanical properties. Analy- simulated body fluid (SBF), containing 6.800 g l1 NaCl, 0.200 g l1
sis and evaluation of the in vitro and in vivo degradation of this Mg– CaCl2, 0.400 g l1 KCl, 0.100 g l1 MgSO4, 2.200 g l1 NaHCO3,
6Zn alloy are also presented in this paper. 0.126 g l1 Na2HPO4 and 0.026 g l1 NaH2PO4 [30]. The pH value
of SBF was adjusted with HCl and NaOH solution to 7.44 and the
2. Materials and methods temperature was maintained at 37 ± 0.5 °C. Pure Mg (>99.99%)
was also tested as a counterpart.
2.1. Materials
2.4.1. Electrochemical measurements
Magnesium alloys with a nominal composition of Mg–6 wt.% Zn Electrochemical measurements were performed using a three
were prepared using high-purity magnesium (P99.99%) and zinc electrode system (PARSTATÒ 2273). A saturated calomel electrode
(P99.9999%) ingots. Melting was carried out at 700–750 °C in a (SCE) and a platinum mesh were used as the reference and counter
high-purity graphite crucible. After about 30 min holding and stir- electrodes, respectively. Potentiodynamic polarization curves were
ring, the melt was cast into a steel mold at about 700 °C. A protec- measured at a scan rate of 1 mV s1. The corrosion rates of the
tive cover gas (99.99% purity argon) was employed throughout the samples were calculated by extrapolating the polarization curve
melting and casting processes. The as cast ingots of Mg–6Zn alloy according to ASTM-G102-89 [31]. Electrochemical impedance
were solid solution treated at about 350 °C for 2 h, followed by spectroscopy (EIS) analysis was also performed at open circuit po-
quenching in water. Finally, the heat-treated alloy sample was tential with a perturbing signal of 5 mV. The frequency varied
hot extruded at about 250 °C with an extrusion ratio of 8:1. from100 to 1 MHz. All of the EIS results were fitted and analyzed
Disk samples for the in vitro degradation and cytotoxicity using ZsimpWin 3.10 Echem software.
experiments, having a diameter of 11.3 mm (so that the cross-sec-
tion area was 1 cm2) and a height of 2.0 mm, were machined from 2.4.2. Immersion tests
the extruded Mg–6Zn rods. Rectangular samples with the dimen- Immersion tests were carried out in conformation with ASTM-
sions 5  5  30 mm were made and used for three-point bending G31-72 [32] (the ratio of surface area to solution volume was
testing. Cylindrical rods with a diameter of 4.5 mm and a length of 1 cm2:20 ml). Samples were removed after 3 and 30 days of
10 mm were used in the animal implant experiments. All samples immersion, rinsed with distilled water and dried at room temper-
were ground with SiC paper up to 1200 grit and polished with ature. Then the surface morphology after immersion was observed
1 lm diamond paste, followed by ultrasonic cleaning in acetone, using scanning electron microscopy (SEM) with an energy disper-
ethanol and distilled water. Before the in vitro cytotoxicity and sive spectrometer (EDS). XRD (Rigaku, D/MAX255) was used to
in vivo degradation experiments all samples were sterilized with determine the phase of corrosion products on the Mg–6Zn surface.
29 kGy of 60Co radiation. Lastly, the samples were cleaned with 180 g l1 chromic acid to re-
move the corrosion products and the degradation rates (in units of
628 S. Zhang et al. / Acta Biomaterialia 6 (2010) 626–640

mm year1) were obtained according to ASTM-G31-72. The corro- hospital. A total of 12 adult New Zealand rabbits (6 females) with
sion rate is given by Eq. (1) [32]: a body weight of 2.0–2.5 kg were used. The rabbits were ran-
domly divided into two groups. In the experimental group one
Corrosion rate ¼ ðK  WÞ  ðA  T  DÞ ð1Þ Mg–6Zn rod was implanted into the femoral shaft of each rabbit.
In the control group each rabbit had a hole drilled in the femoral
where the coefficient K = 8.76  104, W is the weight loss (g), A is
shaft, without any implantation. Before implantation a dose of
the sample area exposed to solution (cm2), T is the exposure time
30 mg kg1 sodium pentobarbital (Shanghai Xinya pharmaceutics
(h) and D is the density of the material (g cm–3).
Ltd.) was administered by intravenous injection. The femoral re-
The pH value of the solution was also recorded during the
gion was cleaned with physiological saline and 75% ethanol. All
immersion tests (PHS-3C pH meter, Lei-ci, Shanghai).
animals received a subcutaneous injection of penicillin as an
anti-inflammatory.
2.5. In vitro loss of mechanical integrity

Magnesium alloys have shown very good properties for orthopedic 2.7.2. Biochemical tests
applications [13]. However, the loss of mechanical integrity during During the experiments 5 ml blood samples were taken from
degradation may inhibit clinical realization [33]. To evaluate the influ- the helix vessel of the rabbits before operation and at 3, 15, 21
ence of in vitro degradation on the mechanical properties rectangular and 28 days after surgery. The blood biochemical tests, including
samples with the dimensions 5  5  30 mm were weighted to an for serum magnesium, serum creatinine (CREA), blood urea nitro-
accuracy of 0.1 mg and then immersed in physiological saline [0.9% gen (BUN), glutamic-pyruvic transaminase (GPT) and creatine
NaCl, Baxter Healthcare (Shanghai) Co. Ltd.]. The samples were taken kinase (CK), were performed with a Hitachi 7600-020 automatic
out, cleaned with ethanol and distilled water, dried in warm air and biochemical analyzer.
weighed each day to monitor the mass change. The immersion tests
continued until the weight loss percentage [weight loss percent- 2.7.3. Radiographic and histological evaluation
age = (Wbefore immersion  Wafter immersion)  Wbefore immersion, where W Radiographs were used to observe the in vivo degradation
means weight] reached about 5%, 10% and 20%, respectively. Finally process during the experiments. Six and 18 weeks post-implanta-
three-point bending tests with a support span of 25 mm were carried tion bone samples around the Mg–6Zn rod were fixed in 4%
out with a MTS 810.22 universal testing machine to observe the influ- formaldehyde and then embedded in methyl methacrylate. Histo-
ence of weight loss on mechanical integrity. The morphology of the logical evaluation was performed on hematoxylin and eosin (HE)
fracture surface was examined by SEM (JEOL JSM 6460). stained sections. The HE stained method was also used to ob-
serve the tissues around the hydrogen gas bubble 3 weeks after
implantation.
2.6. Cytotoxicity assessments
In addition, tetracycline labeling hard tissue fluorescence was
used to observe the interaction between bone and implants. Mean-
L-929 cells were cultured in Dulbecco’s modified Eagle’s med-
while, heart, kidney, spleen and liver tissue from the rabbits were
ium (DMEM) (Gibco), supplemented with 10% fetal bovine serum
also inspected with HE staining to verify whether degradation of
(FBS) in a humidified incubator at 95% relative humidity and 5%
the magnesium alloy harmed these important visceral organs.
CO2 at 37 °C. Cytotoxicity was determined by indirect contact. Ex-
tracts were prepared according to ISO 10993-5: 1999 [34]. The
2.7.4. In vivo degradation characterization
extraction media were serially diluted to 50 and 10% concentration
The rabbits were killed 14 weeks post-implantation. The
after 72 h incubation in a humidified atmosphere with 5% CO2 at
implanted rods were retrieved and weighed and the weight loss
37 °C. The DMEM medium acted as a negative control while DMEM
corrosion rate calculated according to ASTM-G31-72. The surface
medium containing 0.64% phenol as a positive control. Cells were
morphology of the implants was also observed using SEM and EDS.
incubated in 96-well flat bottomed cell culture plates at
2.5  104 cells ml1 medium in each well and incubated for 24 h
2.8. Statistical analysis
to allow cell attachment. Then the medium was replaced by
100 ll extraction medium. After incubation in a humidified atmo-
The two samples t-test was used to determine whether any sig-
sphere for 2, 4 and 7 days cell morphology was observed by optical
nificant difference existed in the cytotoxicity and blood biochemi-
microscopy (Nikon ELWD 0.3 inverted microscope). 3-(4,5-
cal experiments. The statistical significance was defined as P < 0.05.
Dimethylthiazol-2-yl)-2,5-diphenyltetrazolium bromide (MTT)
(Sigma) was then dissolved in phosphate-buffered saline (PBS) at
a concentration of 5 mg ml1. After that 20 ll MTT was added to 3. Results
each well and the samples were incubated for 4 h (37 °C, 5% CO2,
95% relative humidity). Subsequently, 150 ml formazan solution 3.1. Microstructure and mechanical properties
was added to each well and optical density (OD) measurements
were conducted at 490 nm using a Wellscan MK3 spectrophotom- Fig. 1 illustrates the optical microstructure and XRD results for
eter (Labsystem). The cell relative growth rate (RGR) was calcu- the as cast, heat-treated and extruded Mg–6Zn alloys. The micro-
lated according to the following formula: structure of extruded pure Mg is also displayed in Fig. 1d.
It can be seen from Fig. 1a that there were two main phases in
the as cast samples, i.e. the matrix a phase and the second phase c-
RGR ¼ ODtest =ODnegative  100% ð2Þ
MgZn, precipitating along the grain boundary [26]. After solid solu-
tion treatment the c phase disappeared (Fig. 1b) and the alloy had
2.7. In vivo degradation and biocompatibility a supersaturated single phase microstructure. With such a single
phase microstructure the material was much easier to process by
2.7.1. Surgery hot working. The grain size of the extruded samples (Fig. 1c) was
Animal tests were approved by the Ethnics Committee of the much finer than that of the as cast and heat-treated ones, indicat-
6th Hospital of Shanghai Jiao Tong University. The in vivo degrada- ing that hot working had refined the microstructure and thus
tion experiments were performed in the animal laboratory of the should improve the mechanical properties. Neither precipitates
S. Zhang et al. / Acta Biomaterialia 6 (2010) 626–640 629

Fig. 1. Microstructures of (a) as cast, (b) heat-treated, (c) extruded Mg–6Zn alloy and (d) extruded Mg and (e) X-ray diffraction results for the Mg–6Zn alloy.

Table 2
Chemical composition of the Mg–6Zn alloy.

Material Chemical composition (wt.%)


Fe Si Ni Cu Al Mn Zn Mg
Mg–6Zn 0.0038 0.0016 0.0005 0.0005 0.0085 0.0004 5.6210 Balance

nor bulk impurities were observed, suggesting a uniform micro- ple, while after heat treatment and extrusion the c-MgZn peaks
structure in the extruded samples (Fig. 1c). In addition, the grain disappeared in the XRD results, which is consistent with the optical
size of the extruded pure Mg was similar to that of the extruded microstructures in Fig. 1a–c.
Mg–6Zn alloy (Fig. 1d). The XRD results in Fig. 1e also demonstrate The chemical composition of the Mg–6Zn alloy obtained by ICP-
that the c-MgZn peaks can be clearly identified in the as cast sam- AES is listed in Table 2. The impurity content of the Mg–6Zn alloy
630 S. Zhang et al. / Acta Biomaterialia 6 (2010) 626–640

Table 3
Mechanical properties of the extruded Mg–6Zn alloy.

Alloy Modulus (GPa) Yield strength (MPa) Tensile strength (MPa) Elongation (%) Compression strength (MPa)
Mg–Zn 42.3 ± 0.1 169.5 ± 3.6 279.5 ± 2.3 18.8 ± 0.8 433.7 ± 1.4
Mg–Ca [6,17] 239.63 ± 7.21 10.63 ± 0.64 273.2 ± 6.1
Mg–Mn–Zn [38] 44 246 280 20a
a
Deduced from Fig. 4 in Zhang et al. [38].

Fig. 2. Electrochemical measurements (a) polarization curves, (b) Nyquist plots, (c) Bode phase plots and (d) the equivalent circuit for fitting Nyquist plots.

Table 4
In vitro and in vivo degradation rates of Mg and the Mg–6Zn alloy.

Material In vitro (mm year1) In vivo (mm year1)


Electrochemical measurements Immersion tests 14 weeks
3 days 30 days

Mg 0.20 0.43 ± 0.04 0.10 ± 0.07


Mg–6Zn 0.16 0.20 ± 0.05 0.07 ± 0.02 2.32 ± 0.11

was very low, ensuring better degradation properties and tension and compression and higher elongation than the
biocompatibility. extruded Mg–Ca alloy. Although the Mg–6Zn alloy investigated
Table 3 summarizes the mechanical properties of the ex- in this study had a much coarser grain size than that reported
truded Mg–6Zn in comparison with the reported properties of for the Mg–Mn–Ca alloy (>20 lm for the former and <9 lm for
Mg–Ca and Mg–Mn–Zn alloys [6,17]. It should be noted that the latter), both alloys showed comparable mechanical
the Mg–6Zn alloy demonstrated higher ultimate strength in properties.
S. Zhang et al. / Acta Biomaterialia 6 (2010) 626–640 631

Fig. 3. Surface morphology after immersion of (a) Mg for 3 days, (b) Mg–6Zn for 3 days, (c) Mg for 30 days and (d) Mg–Zn for 30 days and (e) the EDS spectrum of the
rectangular area in (b) and (f) EDS of the rectangular area in (d).

3.2. In vitro degradation tests the existence of a protective film on the surface. The corrosion
rates obtained from the polarization curves are listed in Table 4.
3.2.1. Electrochemical measurements It can be seen that the rate of electrochemical degradation of the
Fig. 2 shows typical polarization curves for the Mg–6Zn alloy Mg–6Zn alloy is slower than that of pure Mg. The corrosion poten-
and pure Mg. Current plateaus with different breakdown potentials tial (Ecorr) and breakdown potential (Ept) are also shown in Fig. 2a.
were observed in the anodic parts of the curves (Fig. 2a), indicating The Ecorr value of the Mg–6Zn alloy is greater than that of Mg, indi-
632 S. Zhang et al. / Acta Biomaterialia 6 (2010) 626–640

cating that the Mg–6Zn alloy is less susceptible to corrosion. The


addition of Zn also increased the Ept value (Fig. 2a, Ept1 > Ept2), sug-
gesting that the corrosion layers on the Mg–6Zn sample were more
protective than those on the pure Mg sample.
Fig. 2b and c presents Nyquist plots and Bode phase plots,
respectively. It can be deduced from the Bode phase plots that an
equivalent circuit with two time constants is reasonable. The
equivalent circuit is shown in Fig. 2d, and the polarization resis-
tance Rp (Rp = Rpo + Rct, the higher the Rp, the lower the corrosion
rate [35]) of Mg and Mg–6Zn are shown in Fig. 2b. The Rp of Mg–
6Zn is higher than that of Mg, which is in consistent with the Ecorr
data and electrochemical rates.

3.2.2. Immersion experiments


The degradation rates of the pure Mg and the Mg–6Zn alloy after 3
and 30 days immersion are listed in Table 4. The Mg–6Zn alloy de-
graded more slowly than pure Mg, which is in good agreement with
the electrochemical results. Fig. 3a–d illustrates the surface mor-
phologies and Fig. 3e–f shows the EDS results for the surface corro- Fig. 5. The pH of SBF during 72 h immersion.
sion products on the Mg–6Zn alloy. As shown in Fig. 3a and b, both
pure Mg and the Mg–6Zn alloy sample experienced pitting corrosion
and were covered with partially protective corrosion products.
A number of cracks were observed on the surface of the samples
after 30 days immersion (Fig. 3c and d). The EDS results (Fig. 3e and
f) reveal that the surface corrosion products (rectangular area in
Fig. 3b and d) were rich in O, Mg, P and Ca. The XRD results suggest
that magnesium hydroxide [Mg(OH)2] and hydroxyapatite (HA)
were precipitated on the Mg–6Zn surface (Fig. 4). Furthermore,
Kuwahara et al. [36] pointed out that the corrosion products on
the surface of Mg immersed in Hank’s solution might be amor-
phous (Ca0.86Mg0.14)10(PO4)6(OH)2, a rather complicated com-
pound. In view of the similar ion concentrations in the SBF used
in this study to those in Hank’s solution, there might be some
amorphous phosphates containing magnesium/calcium, as Kuwa-
hara indicated. In fact, a strong background and broadened peaks
can be observed in Fig. 4, which might be due to the presence of
amorphous corrosion products.
Fig. 5 illustrates the pH variation over the first 72 h of the
immersion tests. It can be seen that the pH rose rapidly, i.e. from
7.44 to 8 in 2 h. After 20 h the pH had stabilized. At the end of Fig. 6. The deterioration in bending strength during in vitro degradation.
the immersion tests (after 30 days) the pH was 9.22 for pure Mg
and 9.32 for the Mg–6Zn alloy.
Hydrogen evolution showed a similar trend to the pH value. In
the early stage of immersion both pure Mg and the Mg–6Zn alloy

Fig. 7. The corrosion surface and corrosion holes formed on the samples with 12%
weight loss.

reacted with SBF acutely and a rapid generation of bubbles was ob-
served, indicating a fast rate of hydrogen evolution. After 48 h
Fig. 4. X-ray diffraction pattern of Mg–6Zn immersed in SBF for 30 days. immersion, however, fewer bubbles appeared, suggesting that
S. Zhang et al. / Acta Biomaterialia 6 (2010) 626–640 633

Fig. 8. Morphology of the fracture section after a three-point bending test: (a) extruded Mg–6Zn alloy; (b) after 5.8% weight loss; (c) after 11.9% weight loss; and (d) after 18%
weight loss.

the reaction had slowed down and that the rate of hydrogen evo- healthy, similar to that of the negative control. There was no signif-
lution had decreased. icant difference between the RGR of cells in the extracts and those
It should be noted that the degradation rate measured after in the negative control. According to ISO 10993-5: 1999 [34] the
30 days immersion was lower than that measured after 3 days cytotoxicity of these extracts was Grade 0–1. In other words, the
(Table 4), because the corrosion films, including HA and other phos- Mg–6Zn alloy has a level of biosafety suitable for in cellular
phates, had a protective effect and hence retarded further applications.
degradation.
3.5. In vivo degradation and biocompatibility
3.3. In vitro loss of mechanical integrity
3.5.1. Biochemical tests
Fig. 6 demonstrates the influence of in vitro degradation on the After implantation of the Mg–6Zn rods no rabbits displayed
bending strength of the Mg–6Zn alloy. As shown in Fig. 6, the inflammation and there were no unexpected deaths. The variations
bending strength decreased rapidly in the early stage of degrada- in serum magnesium, CREA, BUN, GPT and CK are shown in Fig. 11.
tion. After that the bending strength deteriorated continuously as There were no significant differences (P > 0.05) in these biochemi-
the percentage weight loss increased. This may be due to surface cal indicators before and after operation, which indicated that deg-
defects such as corrosion holes formed during degradation, as radation of the Mg–6Zn implants did not raise serum Mg2+ levels
shown in Fig. 7. and did not affect kidney and liver function either. The tests indi-
Some dimples were observed on the fracture surface (Fig. 8), cated good biocompatibility of the Mg–6Zn alloy in vivo.
which indicates ductile fracture characteristics of the Mg–6Zn
alloy during the degradation process. The bending tests showed 3.5.2. Viscera histology
that degradation undermined the bending strength, although the Fig. 12 shows HE stained slices of heart, liver, kidney and spleen.
fracture mode was still ductile during degradation. It can be seen that the tissues were normal, which is in good agree-
ment of the biomechanical tests (Fig. 11) and suggests good bio-
3.4. Cytotoxicity assessments compatibility of the Mg–6Zn alloy in vivo.

Fig. 9 shows the morphologies of L-929 cells cultured in differ- 3.5.3. Radiographic evaluation
ent extracts after 7 days incubation. Fig. 10 shows the RGR of L-929 Fig. 13a and b shows the radiographs 3 and 12 weeks post-oper-
cells after 2, 4 and 7 days incubation. It can be seen from Fig. 9 that ation. The Mg–6Zn implant started to degrade in the first 3 weeks,
the cell morphologies in different extracts were normal and as is evident from the observation that the edge of the implant be-
634 S. Zhang et al. / Acta Biomaterialia 6 (2010) 626–640

Fig. 9. L-929 cell morphology after 7 days incubation: (a) negative; (b) 10% extraction; (c) 50% extraction; and (d) 100% extraction.

in the femora and most of the implant had been absorbed


(Fig. 13c).

3.5.4. Bone tissue histology


New bone formation as well as a gap between the residual im-
plant and the surrounding bone tissues can be observed in Fig. 14,
which shows the tetracycline labeling fluorescence results at
14 weeks. In comparison, the femora of rabbits in the control group
had already healed after 3 weeks and the hole had disappeared.
Fig. 15 demonstrates the bone tissue response to the Mg–6Zn
implants 6 and 18 weeks post-implantation. There are newly
formed trabecular and osteoblasts (black arrow in Fig. 15a) at
6 weeks and at 18 weeks more new bone can be found (black ar-
row in Fig. 15b), indicating that the hole in the femur gradually
healed during degradation of the Mg–6Zn alloy.
The tissues around the gas bubble 3 weeks after implantation
are illustrated in Fig. 16, composed of two layers, an inner compact
one and outer loose connective tissue. No obvious inflammatory
response can be observed in Fig. 16. It is hypothesized that the
Fig. 10. RGR of L-929 cells after 2, 4 and 7 days incubation. *P > 0.05. hydrogen gas diffuses into the surrounding tissues [8] and thus
the bubble had disappeared at 6 weeks.
came fuzzy and subcutaneous bubbles appeared (Fig. 13a). No ad-
verse effects due to these gas bubbles were observed in the rabbits 3.5.5. In vivo degradation characterization
and they disappeared after 6 weeks without any intervention, After the rabbits were killed 14 weeks post-implantation the
which is in agreement with the literature [4,6]. After 12 weeks residual Mg–6Zn rod material was weighed. It was found that
the implant was too blurry to be recognized (Fig. 13b). The radio- the residual weight was only 13% of the original weight, i.e. 87%
graphs offer evidence that the implant gradually degraded within of the implant had been degraded. The in vivo degradation rate ob-
the bone. In fact, when the rabbits were killed 14 weeks after served in this study was faster than in previous reports. For exam-
implantation an obvious irregular shaped hole could be observed ple, a Mg–Mn–Zn alloy showed 54% degradation 18 weeks post-
S. Zhang et al. / Acta Biomaterialia 6 (2010) 626–640 635

Fig. 11. Blood biochemical indicators: (a) serum magnesium; (b) CREA; (c) BUN; (d) GPT; and (e) CK. The blue dashed lines in (a–e) represent the level before surgery.

operation [16]. The in vivo corrosion rate according to Eq. (1) was 4. Discussion
2.32 mm year1 (also listed in Table 4).
The surface morphology of the residual rod material 14 weeks 4.1. Mechanical properties
after implantation is shown in Fig. 17a, showing a rough surface
covered with a thick layer of corrosion products. The corrosion Zinc is a common alloying element for magnesium and has a
products contained large amounts of O, Mg, P, Ca, Na and Zn solution hardening effect on magnesium alloys [25]. Although
(Fig. 17b). It can be seen that the in vivo degradation morphology the maximum solubility drops to 1.6 wt.% (i.e. 0.6 at.%) at
is different from the in vitro morphology and the compositions are 25 °C in the equilibrium state [26], a solid solution treatment
also slightly different, suggesting dissimilar degradation behaviors can be carried out to obtain a supersaturated solution to avoid
of this magnesium alloy in vitro and in vivo. precipitation of the c-MgZn phase. In this way a uniform micro-
636 S. Zhang et al. / Acta Biomaterialia 6 (2010) 626–640

Fig. 12. HE stained slices of (a) heart, (b) liver, (c) kidney and (d) spleen. Scale bar 100 lm.

structure can be obtained (Fig. 1) and the corrosion process can The existence of chloride ions (Cl) transforms Mg(OH)2 into
be slowed, due to the homogeneity. Moreover, the mechanical soluble MgCl2 [6], resulting in excess OH ions in the solution.
properties can also be enhanced by solution hardening. For Eventually the pH will rise (shown in Fig. 5). In fact, even though
example, Table 3 suggests that zinc can effectively improve the the bulk solution has a pH value as low as 4, the local pH near
mechanical properties. the surface of the Mg could be >10 [39]. As a result, if the solution
Hot working (hot extrusion in this paper) can also refine the contains ions such as PO43, Ca2+, etc. HA [Ca10(PO4)6(OH)2] is
grain size and hence increase the yield strength and tensile likely to nucleate and grow on the magnesium surface due to the
strength according to the Hall–Petch relationship [11]. Compared supersaturated condition at high pH [40]. This phenomenon ex-
with degradable polymeric materials such as l-HA/PLLA 50/50 plains the detection of HA by XRD in this study (Fig. 4).
(tensile strength 103 MPa, Young’s modulus 12.3 GPa) [37], the Moreover, when Mg2+ ions dissolve into the solution, phos-
Mg–6Zn alloy has improved mechanical properties. The elastic phates containing Mg/Ca form and tightly attach to the matrix. In
modulus of the magnesium alloy in this paper (42.3 GPa) is also general, the reaction among HnPO4(3–n)– (representing various
closer to that of human femur bone (15–20 GPa [38]) than that phosphate ions, where n = 0, 1 or 2), Ca2+ and Mg2+ could be de-
of polymers and titanium alloys. scribed primarily as:
However, the mechanical properties will deteriorate gradu-
þ Ca2þ þ Mg2þ þ OH ! Mgx Cay ðPO4 Þz
ð3nÞ
ally during the degradation process, as shown in Fig. 6. In this re- Hn PO4
gard, the degradation process must be precisely controlled so ðinsoluble productsÞ ð6Þ
that the mechanical integrity meets requirements before the bone
has healed. Nevertheless, long-term in vivo deterioration of Taking the excess OH into consideration, some complicated
mechanical integrity needs to be investigated in detail. compounds [represented by MgxCay(PO4)z(OH)] might precipitate
on the surface (Fig. 3). Previous studies [4,16] have shown that
the corrosion layer containing such magnesium-substituted cal-
4.2. The degradation process
cium phosphate compounds on Mg can promote osteoinductivity
and osteoconductivity, predicting good biocompatibility of
Magnesium in aqueous solution (for instance SBF) dissolves
magnesium.
according to the following equations [1]:
It is difficult to determine the in vivo degradation rate precisely
anodic reaction : Mg ! Mg2þ þ 2e ð3Þ because the circumstances in vivo are quite complicated. The Mg–
cathodic reaction : 2H2 O þ 2e ! H2 þ 2OH 
ð4Þ 6Zn implant had been degraded by 87% 14 weeks post-operation
and the corrosion rate was 2.32 mm year1 according to Eq. (1).
Mg2þ þ 2OH ! MgðOHÞ2 ð5Þ Compared with previous investigations (Table 5), it seems that
S. Zhang et al. / Acta Biomaterialia 6 (2010) 626–640 637

Fig. 13. Radiographs (a) 3 weeks post-operation, (b) 12 weeks post-operation and (c) macroscopic photograph 14 weeks post-operation, with an irregular shaped hole in the
femur (arrow).

the Mg–6Zn in the present study degraded slightly faster. Several


methods can be used to regulate the degradation rate, for instance
calcium phosphate coating and anodic oxidation.
Zinc is able to elevate the corrosion potential of magnesium alloys
and improve the corrosion resistance [41], which is in agreement
with this study (Table 4). In addition, zinc has beneficial effects on
the corrosion film and can reduce the effect of impurities once the
tolerance limits have been exceeded [42]. Furthermore, it has been
found that zinc can also elevate the charge transfer resistance of
magnesium and thus reduce the corrosion rate [43]. The degradation
rates of the Mg–6Zn alloy (0.20 ± 0.05 for 3 days and 0.07 ± 0.02 for
30 days) are slower than those of pure Mg (0.43 ± 0.04 for 3 days and
0.10 ± 0.07 for 30 days).
For both the Mg–6Zn alloy and pure Mg the degradation
rates for 30 days were lower than those for 3 days, owing to
the protective layer on the surface (shown in Fig. 3a–d).
During the early stage of immersion in SBF Mg and the Mg–
Fig. 14. Tetracycline labeling 14 weeks post-operation. 6Zn alloy degraded quickly, accompanied by the rapid formation
638 S. Zhang et al. / Acta Biomaterialia 6 (2010) 626–640

Fig. 15. HE stained bone surrounding Mg–Zn rods: 6 weeks (a) and 18 weeks (b) post-implantation. Scale bar 500 lm.

Fig. 16. HE stained tissues around the gas bubble 3 weeks after implantation.

of an insoluble protective corrosion layer, which retarded


degradation.
It should be noted that the compositions and the morphologies
of the in vitro and in vivo degradation products are different. This
is mainly due to the different corrosive environments. The in vitro
testing solution was SBF, containing inorganic ions such as Cl,
H2PO4 and Ca2+. Thus, the corrosion products were mineral-like
and the corrosion layer could be clearly observed, which included
HA, Mg(OH)2 and other amorphous magnesium-substituted cal-
cium phosphates. However, there were numbers of organic compo-
nents in the in vivo environment, such as proteins and cells. Rettig
et al. found that albumin influenced the corrosion process of mag-
nesium alloys in SBF [44]. As a consequence, the in vivo degrada-
tion products are quite complex. The composition of the in vivo
degradation products in Fig. 17b shows a high nitrogen level in
the corrosion layer, indicating the possible adsorption and/or adhe-
sion of proteins, cells or other tissue fragments to the magnesium
Fig. 17. Surface morphology and composition of the corrosion products of the
alloy.
implant after 14 weeks degradation: (a) SEM; and (b) EDS of framed area in (a).
The different morphologies and compositions in vitro and
in vivo suggest that degradation of the magnesium alloy will be
greatly influenced by the surrounding environments. Thus, if the
magnesium alloy is used in different parts of the body (e.g. in fe- behaviors must be taken into consideration, otherwise the rate of
mur marrow or within muscle) possible different degradation degradation may be estimate incorrectly.
S. Zhang et al. / Acta Biomaterialia 6 (2010) 626–640 639

Table 5 to adjust the degradation rate, surface modification technologies


Comparison of in vivo degradation rates. (e.g. electrodeposition of HA coating, spray coating with PLGA,
Percent Degradation rate Reference etc.) are presently being applied to the Mg–6Zn alloy by our group.
degradation (mm year1)
Mg–6Zn 87% after 14 weeks 2.32 Present study
Mg–Ca 1.27a [6] 5. Conclusions
AZ91D 3.516  104 [5]
LAE442 1.205  104 [5]
In this paper a binary Mg–6 wt.% Zn alloy was investigated as a
Mg–Mn–Zn 54% after 18 weeks [16]
biomedical degradable magnesium. The following conclusions can
a
The Mg–Ca degradation rate was deduced from Li et al. [6] be drawn.
(2.28 mg mm2 year1).
The Mg–6Zn alloy fabricated with high-purity raw materials
and using a clean melting process had very low contents of impu-
4.3. Biocompatibility rities. After solid solution treatment and hot working the grain size
of Mg–6Zn alloy was refined and a uniform single phase was
The in vitro cytotoxicity of Mg–6Zn was found to be Grade 0–1, obtained.
indicating that the Mg–6Zn alloy is safe as an implantable material. The mechanical properties of the Mg–6Zn alloy were suitable
There was a gap between the bone and the residual implant for implant applications, i.e. the tensile strength and elongation
14 weeks post-operation (Fig. 14). This may be due to the rapid achieved were 279.5 MPa and 18.8%, respectively. Degradation
degradation and rapidly elevated ambient pH. As a consequence, of the Mg–6Zn alloy resulted in a deterioration in the mechanical
osteoblasts were unable to proliferate and adhere very well. Never- properties and mechanical integrity was rapidly lost in the early
theless, there was still some newly formed bone surrounding the stages of corrosion, but more slowly in the later stages of in vitro
implant despite the fast degradation, and the amount of new bone degradation.
increased from 6 to 18 weeks post-operation (Fig. 15). Zreiqat et al. Zinc elevated the corrosion potential of the magnesium alloy
[45] found that magnesium ions could enhance the adhesion of hu- and the in vitro degradation rate of the Mg–6Zn alloy was slower
man bone-derived cells (HBDC) and increased the levels of a5b1- than that of high-purity magnesium in SBF. A protective layer of
and b1-integrin receptors. It has been reported that zinc can also HA and other Mg/Ca phosphates formed on the surface of Mg–
promote bone formation [46]. Therefore, it is postulated that the 6Zn when immersed in SBF.
Mg–6Zn alloy has good biocompatibility in bone. In vitro cytotoxicity to L-929 cells implies that the Mg–6Zn alloy
In addition, the corrosion layer on the Mg–6Zn alloy contained is safe for cellular applications, with a cytotoxicity grade of 0–1.
Mg, Ca and P, which could reinforce osteoblast activity. Hence, we Animal implant experiments indicated that the Mg–6Zn alloy
come to the conclusion that rapid degradation of the Mg–6Zn alloy gradually degraded within the femoral shaft, with a degradation
was not harmful to adjacent bone tissues. A previous study [9] also rate of 2.32 mm year1. Subcutaneous hydrogen gas bubbles pro-
confirmed that fast degrading AZ91D scaffolds induced extended duced by degradation disappeared 6 weeks after implantation
peri-implant bone remodeling with good biocompatibility. without discernable adverse effects. It was found that the bubble
The excess magnesium produced by degradation can be ex- was enveloped by two layers of connective tissue. Owing to the rel-
creted by the kidneys, as indicated in Fig. 11a, with no kidney, liver atively rapid degradation, a gap between the implant and sur-
or heart disorders being observed during degradation (Figs. 11 and rounding bone tissues was observed, however, there was still
12), and no adverse effects of bubbles were observed. These results newly formed trabeculae and osteoblasts.
also imply good in vivo biocompatibility of the Mg–6Zn alloy. According to the biochemical indicators and the HE stained sec-
Nevertheless, the mechanism of hydrogen absorption and tions there were no disorders of the heart, kidney, liver and spleen
whether the hydrogen can be metabolized or will accumulate in cer- and no negative effects of zinc release were observed, indicating
tain organs are still unknown and need further detailed investiga- that the binary Mg–6Zn alloy had good biocompatibility in vivo.
tion. In addition, it can be seen from Fig. 16 that there are dual
connective tissues around the bubble, which may affect the adhesion Acknowledgements
of cells and be one reason for the gap between the residual implant
and the surrounding bone, as shown in Fig. 14. If the degradation rate The authors are grateful for the supports from the National Nat-
is slow enough it is possible to avoid bubble formation, as the hydro- ural Science Foundation of China (No. 30772182), Shanghai Jiao
gen can diffuse into the surrounding tissues [8]. Tong University Interdisciplinary Research Grants (Grant No.
Zinc is an essential element for humans, is absolutely necessary YG2007MS26 and YG2009MS53) and the 863 High-tech Plan of
and is non-toxic except at extreme exposures. The human require- China (No.2009AA03Z424).
ment for zinc is estimated to be 15 mg day1 [23]. The degradation
rate of Mg–6Zn within rabbit femora in this study was
2.32 mm year1, i.e. 1.14 mg cm2 day1. The zinc content of Appendix A. Figures with essential colour discrimination
the Mg–6Zn alloy in this study was 5.621 wt.% (Table 2), so the zinc
release rate will be 0.0641 mg cm2 day1. For an implant with Certain figures in this article, particularly Figs. 1–7 and 11–17, are
the dimensions 4.5  10 mm (cylindrical, with a surface area of difficult to interpret in black and white. The full colour images can be
1.732 cm2), the zinc release rate would be about 0.11 mg day1, found in the on-line version, at doi:10.1016/j.actbio.2009.06.028.
which is much lower than the suggested intake of 15 mg day1
for adults. Besides, the released zinc could be absorbed by the sur-
rounding tissues [16] and excreted through the gastrointestinal References
route and the kidney [23]. Therefore, it is proposed that the zinc re-
[1] Song G, Atrens A, StJohn D, Nairn J, Li Y. The electrochemical corrosion of pure
lease during degradation is safe. magnesium in 1 N NaCl. Corros Sci 1997;39(5):855–75.
In summary, even though the Mg–6Zn alloy degraded rapidly [2] Song G, Atrens A, StJohn D, Wu X, Nairn J. The anodic dissolution of magnesium
in chloride and sulphate solutions. Corros Sci 1997;39(10–11):1981–2004.
in vivo, there was still newly formed bone surrounding the im-
[3] Heublein B, Rohde R, Kaese V, Niemeyer M, Hartung W, Haverich A.
plants and no disorders were found in heart, liver, kidney and Biocorrosion of magnesium alloys: a new principle in cardiovascular implant
spleen, suggesting good biocompatibility. Furthermore, in order technology? Heart 2003;89(6):651–6.
640 S. Zhang et al. / Acta Biomaterialia 6 (2010) 626–640

[4] Witte F, Kaese V, Haferkamp H, Switzer E, Meyer-Lindenberg A, Wirth CJ, et al. [27] Zhang X. In vivo biodegradable binary Mg–Zn alloy. Chinese Patent no. ZL
In vivo corrosion of four magnesium alloys and the associated bone response. 200510111795.4.
Biomaterials 2005;26(17):3557–63. [28] American Society for Testing and Materials. ASTM-E8-04: standard test methods
[5] Witte F, Fischer J, Nellesen J, Crostack H-A, Kaese V, Pisch A, et al. In vitro and for tension testing of metallic materials. In: Annual book of ASTM standards.
in vivo corrosion measurements of magnesium alloys. Biomaterials Philadelphia, PA, USA: American Society for Testing and Materials; 2004.
2006;27(7):1013–8. [29] American Society for Testing and Materials. ASTM E9-89a (2000): standard
[6] Li Z, Gu X, Lou S, Zheng Y. The development of binary Mg–Ca alloys for use as test methods of compression testing of metallic materials at room
biodegradable materials within bone. Biomaterials 2008;29(10):1329–44. temperature. In: Annual Book of ASTM Standards. Philadelphia, PA, USA:
[7] Kim WC, Kim JG, Lee JY, Seok HK. Influence of Ca on the corrosion properties of American Society for Testing and Materials; 2000.
magnesium for biomaterials. Mater Lett 2008;62(25):4146–8. [30] Standardization Administration of the PRC. GB/T 16886.15-2003: Biological
[8] Witte F, Ulrich H, Rudert M, Willbold E. Biodegradable magnesium scaffolds. Evaluation of Medical Devices. Part 15. Identification and Quantification of
Part I. Appropriate inflammatory response. J Biomed Mater Res A Degradation Products from Metals and Alloys.
2007;81A(3):748–56. [31] American Society for Testing and Materials. ASTM-G102-89: standard practice
[9] Witte F, Ulrich H, Palm C, Willbold E. Biodegradable magnesium scaffolds. Part for calculation for corrosion rates and related information from
II. Peri-implant bone remodeling. J Biomed Mater Res A 2007;81A(3):757–65. electrochemical measurements. In: Annual Book of ASTM Standards.
[10] Witte F, Feyerabend F, Maier P, Fischer J, Stormer M, Blawert C, et al. Philadelphia, PA, USA: American Society for Testing and Materials; 1999.
Biodegradable magnesium–hydroxyapatite metal matrix composites. [32] American Society for Testing and Materials. ASTM-G31-72: standard
Biomaterials 2007;28(13):2163–74. practice for laboratory immersion corrosion testing of metals. In: Annual
[11] Zhang E, He W, Du H, Yang K. Microstructure, mechanical properties and Book of ASTM Standards. Philadelphia, PA: American Society for Testing and
corrosion properties of Mg–Zn–Y alloys with low Zn content. Mater Sci Eng A Materials; 2004.
2008;488(1–2):102–11. [33] Kannan MB, Raman RKS. In vitro degradation and mechanical integrity of
[12] Song G. Control of biodegradation of biocompatible magnesium alloys. Corros calcium-containing magnesium alloys in modified-simulated body fluid.
Sci 2007;49(4):1696–701. Biomaterials 2008;29(15):2306–14.
[13] Staiger MP, Pietak AM, Huadmai J, Dias G. Magnesium and its alloys as [34] ANSI/AAMI. ISO 10993-5: 1999. Biological evaluation of medical devices. Part
orthopedic biomaterials: a review. Biomaterials 2006;27(9):1728–34. 5. Tests for cytotoxicity: in vitro methods. Arlington, VA: ANSI/AAMI.
[14] Levesque J, Hermawan H, Dub D, Mantovani D. Design of a pseudo- [35] Rondellia G, Torricellib P, Finib M, Giardinob R. In vitro corrosion study by EIS
physiological test bench specific to the development of biodegradable of a nickel-free stainless steel for orthopaedic applications. Biomaterials
metallic biomaterials. Acta Biomater 2008;4(2):284–95. 2005;26:739–44.
[15] Song GL, Song SZ. A possible biodegradable magnesium implant material. Adv [36] Kuwahara H, Al-Abdullat Y, Mazaki N, Tsutsumi S, Aizawa T. Precipitation of
Eng Mater 2007;9(4):298–302. magnesium apatite on pure magnesium surface during immersing in Hank’s
[16] Xu LP, Yu GN, Zhang E, Pan F, Yang K. In vivo corrosion behavior of Mg–Mn–Zn solution. Mater Trans 2001;42(7):1317–21.
alloy for bone implant application. J Biomed Mater Res A 2007;83A(3):703–11. [37] Shikinami Y, Okuno M. Bioresorbable devices made of forged composites of
[17] Wan Y, Xiong G, Luo H, He F, Huang Y, Zhou X. Preparation and hydroxyapatite (HA) particles and poly-L-lactide (PLLA). Part I. Basic
characterization of a new biomedical magnesium–calcium alloy. Mater characteristics. Biomaterials 1999;20:859–77.
Design 2008;29(10):2034–7. [38] Zhang E, Yin D, Xu L, Yang L, Yang K. Microstructure, mechanical and corrosion
[18] Zartner P, Cesnjevar R, Singer H, Weyand M. First successful implantation of a properties and biocompatibility of Mg–Zn–Mn alloys for biomedical
biodegradable metal stent into the left pulmonary artery of a preterm baby. application. Mater Sci Eng C. doi:10.1016/j.msec.2008.08.024.
Catheter Cardiovasc Interv 2005;66(4):590–4. [39] Simaranov A, Sokolova I, Marshakov A, Mikhailovskii Y. Corrosion-
[19] El-Rahman SSA. Neuropathology of aluminum toxicity in rats (glutamate and electrochemical behavior of magnesium in acidic media, containing
GABA impairment). Pharmacol Res 2003;47(3):189–94. oxidants. Prot Metal 1991;27(3):329–34.
[20] Ku C-H, Pioletti DP, Browne M, Gregson PJ. Effect of different Ti–6Al–4V surface [40] Jonasova L, Muller F, Helebrant A, Strnad J, Greil P. Biomimetic apatite
treatments on osteoblasts behaviour. Biomaterials 2002;23(6):1447–54. formation on chemically treated titanium. Biomaterials 2004;25:1187–94.
[21] Yumiko N, Yukari T, Yasuhide T, Tadashi S, Yoshio I. Differences in behavior [41] Shi Z, Song G, Atrens A. Corrosion resistance of anodised single-phase Mg
among the chlorides of seven rare earth elements administered intravenously alloys. Surf Coat Technol 2006;201:492–503.
to rats. Fundam Appl Toxicol 1997;37:106–16. [42] Song GL, Atrens A. Corrosion mechanisms of magnesium alloys. Adv Eng Mater
[22] Yang W, Zhang P, Liu J, Xue Y. Effect of long-term intake of Y3+ in drinking water on 1999;1(1):11–33.
gene expression in brains of rats. J Rare Earth 2006;24(3):369–73. [43] Zhang SX, Li JN, Song Y, Zhao CL, Zhang XN, Xie CY, et al. In vitro degradation,
[23] Tapiero H, Tew KD. Trace elements in human physiology and pathology: zinc hemolysis and MC3T3-E1 cell adhesion of biodegradable Mg–Zn alloy. Mater
and metallothioneins. Biomed Pharmacother 2003;57(9):399–411. Sci Eng C 2009. doi:10.1016/j.msec.2009.03.001.
[24] Haferkamp H, Bach F-W, Kaese V, Möhwald K, Niemeyer M, Schreckenberger [44] Rettig R, Virtanen S. Time-dependent electrochemical characterization of the
H, et al. Magnesium corrosion-processes, protection of anode and cathode. In: corrosion of a magnesium rare-earth alloy in simulated body fluids. J Biomed
Kainer KU, editor. Magnesium-alloys and technology. Weinheim: Wiley-VCH; Mater Res A 2008;85A(1):167–75.
2003. p. 226–7. [45] Zreiqat H, Howlett C, Zannettino A, Evans P, Schulze-Tanzil G, Knabe C, et al.
[25] Mordike BL, Lukác P. Physical metallurgy. In: Friedrich HE, Mordike BL, editors. Mechanisms of magnesium-stimulated adhesion of osteoblastic cells to
Magnesium technology – metallurgy, design data, applications. Berlin: Springer; commonly used orthopaedic implants. J Biomed Mater Res A 2002;62:175–84.
2006. p. 76–7. [46] Hashizume M, Yamaguchi M. Stimulatory effect of beta-alanyl-L-histidinato
[26] Okamoto H. Comment on Mg–Zn (magnesium–zinc). J Phase Equilibria Diffus zinc on cell proliferation is dependent on protein synthesis in osteoblastic
1994;15(1):129–30. MC3T3–E1 cells. Mol Cell Biochem 1993;122(1):59–64.

You might also like