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

Acta Biomaterialia 64 (2017) 421–436

Contents lists available at ScienceDirect

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

Full length article

Development of magnesium-based biodegradable metals with dietary


trace element germanium as orthopaedic implant applications
Dong Bian a, Weirui Zhou a, Jiuxu Deng b, Yang Liu a, Wenting Li a, Xiao Chu c, Peng Xiu d, Hong Cai d,
Yuhui Kou b, Baoguo Jiang b,⇑⇑, Yufeng Zheng a,⇑
a
Department of Materials Science and Engineering, College of Engineering, Peking University, Beijing 100871, China
b
Department of Trauma and Orthopedics, People’s Hospital Peking University, Beijing 100044, China
c
Department of Orthopedics, Guangdong Key Lab of Orthopaedic Technology and Implant Materials, Guangzhou General Hospital of Guangzhou Military Command,
Guangzhou 510010, China
d
Department of Orthopedics, Peking University Third Hospital, Beijing 100191, China

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

Article history: From the perspective of element biosafety and dietetics, the ideal alloying elements for magnesium
Received 19 June 2017 should be those which are essential to or naturally presented in human body. Element germanium is a
Received in revised form 13 September unique metalloid in the carbon group, chemically similar to its group neighbors, Si and Sn. It is a dietary
2017
trace element that naturally presents in human body. Physiological role of Ge is still unanswered, but it
Accepted 3 October 2017
Available online 4 October 2017
might be necessary to ensure normal functioning of the body. In present study, novel magnesium alloys
with dietary trace element Ge were developed. Feasibility of those alloys to be used as orthopaedic
implant applications was systematically evaluated. Mg-Ge alloys consisted of a-Mg matrix and eutectic
Keywords:
Mg-Ge alloy
phases (a-Mg + Mg2Ge). Mechanical properties of Mg-Ge alloys were comparable to current Mg-Ca,
Biodegradable metals Mg-Zn and Mg-Sr biodegradable metals. As-rolled Mg-3Ge alloy exhibited outstanding corrosion resis-
Degradation tance in vitro (0.02 mm/y, electrochemical) with decent corrosion rate in vivo (0.6 mm/y, in rabbit tibia).
In vivo New bone could directly lay down onto the implant and grew along its surface. After 3 months, bone and
Osseointegration implant were closely integrated, indicating well osseointegration being obtained. Generally, this is a
pioneering study on the in vitro and in vivo performances of novel Mg-Ge based biodegradable metals,
and will benefit the future development of this alloy system.

Statement of Significance

The ideal alloying elements for magnesium-based biodegradable metals should be those which are essen-
tial to or naturally presented in human body. Element germanium is a unique metalloid in the carbon
group. It is a dietary trace element that naturally presents in human body. In present study, feasibility
of Mg-Ge alloys to be utilized as orthopedic applications was systematically investigated, mainly focusing
on the microstructure, mechanical property, corrosion behavior and biocompatibility. Our findings
showed that Mg-3Ge alloy exhibited superior corrosion resistance to current Mg-Ca, Mg-Zn and Mg-Sr
alloys with favorable biocompatibility. This is a pioneering study on the in vitro & in vivo performances
of Mg-Ge biodegradable metals, and will benefit the future development of this alloy system.
Ó 2017 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.

1. Introduction attracted great attention since the early 2000 s [1–4]. The ideal
Mg-based biodegradable metals are expected to corrode gradu-
Crystalline magnesium (Mg) and magnesium-based alloys ally in vivo, with an appropriate host response, then dissolve
specifically designed for biodegradable implant applications have completely upon fulfilling the mission to assist with the tissue
healing [5]. During degradation, elements in the implant matrix
will release in the form of metal ions and then they participate
⇑ Corresponding author.
in tissue healing positively, neutrally or negatively, depending on
⇑⇑ Co-corresponding author.
the dose and releasing rate. Therefore, besides magnesium, the
E-mail addresses: jiangbaoguo@vip.sina.com (B. Jiang), yfzheng@pku.edu.cn
(Y. Zheng). major components of Mg-based biodegradable metals should

https://doi.org/10.1016/j.actbio.2017.10.004
1742-7061/Ó 2017 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.
422 D. Bian et al. / Acta Biomaterialia 64 (2017) 421–436

be preferably essential elements or trace elements in human 2. Experimental details


body.
Many kinds of essential/possibly essential elements found in 2.1. Material preparation
human body could be used as alloying elements to improve
in vitro and in vivo performances of Mg-based biomaterials, includ- Mg-Ge alloys were melted and casted by using commercial
ing corrosion control, mechanical improvement, biological regula- magnesium (99.95 wt%) and commercial pure Ge powder
tion, etc. Till now, many kinds of biomedical Mg-based alloys with (99.9 wt%) under the protection of a mixed gas atmosphere of
essential/possibly essential elements addition have been devel- SF6 (1 vol%) and CO2 (99 vol%). Briefly, Mg and Ge powders were
oped, including binary Mg-Ca [6–8], Mg-Zn [9,10], Mg-Mn [11], well blended in a mixer, then melted at 700–800 °C for 30 min
Mg-Fe [12], Mg-Cu [13,14], Mg-Sr [15–18], Mg-Si [19], Mg-Sn before casting. The ingots were machined into 5 mm thick plates
[20–23], Mg-Li [24] and multicomponent Mg alloys based on these and homogenized at 400 °C for 3 h before rolling. Afterwards, they
binary alloys, Mg-Zn-Ca [25] and Mg-Si-Ca-Zn [26], etc. were hot rolled into thin plates with a final thickness of 2.5 mm at
Germanium (Ge) is a lustrous, hard, grayish-white metalloid in a rolling reduction of 0.2 mm each pass. The plate was annealed at
the carbon group, chemically similar to its group neighbors, Si and 400 °C for 10 min at each rolling interval. The actual Ge contents in
Sn. It is usually used as a semiconductor in transistors and various the Mg-xGe alloys were 1.38 ± 0.30, 2.53 ± 0.21 and 3.26 ± 0.20 wt
other electronic devices. Ge presents in human body and it is a %, respectively. Specimens for microstructural characterization,
dietary trace element which might be necessary to ensure normal corrosion test and in vitro biocompatibility evaluation were cut
functioning of the body [27]. For the general population, the major into 10  10  2 mm3 disks. Cylindrical pins with a diameter of
source of Ge is from food. Ge presents in practically all food and 2.2 mm and length of 10 mm were machined parallel to the rolling
only in minute amounts [28]. The estimated average dietary intake direction for in vivo test. Samples were mechanically grounded
of Ge in humans is in the range of 0.4–3.4 mg per day. Ge normal- with SiC sandpaper up to 2000 grit, ultrasonically cleaned in ace-
izes many physiological functions such as lowering of high blood tone and absolute ethanol, then dried in hot air. For cell assays,
pressure in humans, restoring deviant blood characteristics to their samples were sterilized under ultraviolet radiation for 2 h before
normal range including pH, glucose, the minerals sodium, potas- use. Pins for animal test were separately encapsulated and steril-
sium, etc. [29]. Many pharmacological effects of some organic ger- ized under Co60 c ray radiation at 25 KGy.
manium compounds have been investigated. They include
antitumor, anticancer, anti-inflammation, antioxidation, radiopro-
2.2. Microstructural characterization
tection and pro-immune functions [27]. Organic germanium
appears to be remarkably safe while relatively high doses of ger-
Specimens for microstructural observation were polished into
manium dioxide and other inorganic germanium compounds can
mirror-like surface with 5 lm diamond polishing paste. After-
cause severe poisoning including fatal cases [28,30]. Animal stud-
wards, they were etched in 4% HNO3/alcohol solution and observed
ies have already showed promising results, despite sufficient clin-
under an optical microscope (Olympus BX51M, Japan). Average
ical trials are still needed [28].
grain size and phase proportion were calculated in ImageJ software
From the viewpoint of material science, limited information
(ImageJ 1.43 u, USA) by analyzing microstructural images (n = 5).
about Mg-Ge alloys are related to thermodynamic characters
An X-ray diffractometer (XRD, Rigaku DMAX 2400, Japan) was
[31,32], and the effects of Ge on the microstructure and mechan-
employed to identify the constituent phases by using Cu Ka radia-
ical property [33,34]. To the authors’ knowledge, no report on
tion at a scan rate of 4°/min operated at 40 kV and 100 mA.
the corrosion behavior, in vitro or in vivo biocompatibility of
Mg-Ge alloys can be found. Recently, Liu et al. [33] reported
their experimental results on the binary Mg-Ge alloys. They 2.3. Tensile test
found that microstructures of Mg-Ge binary alloys consisted of
two components: a-Mg dendrites and a-Mg + Mg2Ge eutectic Tensile specimens were machined parallel to the rolling direc-
phase. Grain size of a-Mg dendrites was effectively refined with tion according to ASTM-E8-04 [36]. Tensile tests were performed
Ge addition and tensile strengths of Mg-Ge binary alloys were at a crosshead speed of 1 mm/min on a universal material testing
enhanced while elongations were reduced. Benefits of Ge on machine (Instron 5969, USA) at room temperature. An average of
thermal stability and mechanical property of Mg have also been at least three parallel samples was taken for each group.
verified in another work [34]. Even though no referable literature
of Ge on Mg corrosion can be found, effects of this intrinsic met- 2.4. Electrochemical test
alloid/semimetallic element on Mg degradation is really interest-
ing. The continuously net-like second phase (a-Mg + Mg2Ge) in A traditional three-electrode cell system was used for electro-
Mg-Ge alloys, as reported in previous work [33,34], might act chemical test with a platinum foil as counter electrode (CE), a sat-
as a barrier during corrosion [35]. urated calomel electrode (SCE) as reference electrode.
In the present study, binary Mg-xGe (x = 1.5, 2.5 and 3 wt%) Experimental samples with an exposed area of 1 cm2 acted as
alloys were designed and fabricated through traditional casting. working electrode. Measurements were carried out in Hank’s solu-
The Ge content range (1.5–3 wt%) was determined by referring to tion [37], on an electrochemical workstation (CHI660C, China).
Refs. [33] and [34], as the strength was not further improved when Open circuit potential (OCP) was continuously monitored for
Ge content exceeded 3.4 wt% while significantly impaired plastic- 3600 s and then electrochemical impedance spectroscopy (EIS)
ity occurred. For biosafety purpose, Ge content should also be lim- was measured from 100 kHz to 10 mHz using a 5 mV amplitude
ited in a proper low range. To further improve mechanical perturbation. Potentiodynamic polarization tests were performed
property, these binary alloys were hot rolled into thin plates. Their at a scanning rate of 1 mV/s and the initial potential was 300 mV
feasibility to be utilized as orthopedic applications was systemati- below the corrosion potential (Ecorr). Electrochemical parameters
cally investigated, mainly focusing on the microstructure, mechan- and corrosion rates were calculated according to ASTM-G102-89
ical property, corrosion behavior, in vitro cytocompatibility and [38]. Three duplicate samples were taken for each material for sta-
in vivo biocompatibility. tistical analysis.
D. Bian et al. / Acta Biomaterialia 64 (2017) 421–436 423

2.5. Immersion test at 37 °C. Afterwards, 150 lL of chromogenic agent (potassium fer-
ricyanide) was added into the mixed solution. The spectrophoto-
Immersion tests were carried out in Hank’s solution at metric absorbance was measured at 520 nm by a microplate
37 ± 1 °C according to ASTM-G31-72 [39], with an exposure ratio reader (Bio-RAD680). For normalization, total protein content
of 20 mL/cm2. The volume of emerged hydrogen was recorded by was determined using a Bicinchoninic Acid (BCA) protein assay
using a calibrated burette in accordance with Ref. [35]. The corre- kit (Beijing Biosea Biotechnology, China). The ALP activity was
sponding corrosion rate was calculated based on the following expressed and normalized by the total protein content (U/gprot).
equation [40]: pH = 2.279VH (PH: corrosion rate, mm/y; VH: hydro-
gen evolution rate, mL/cm2/d). After immersion for 3 and 20 days,
samples were removed out of the solution, gently rinsed with 2.8. Animal test
distilled water, and dried in open air. Surface morphology as well
as cross section morphology of the corroded samples was observed 2.8.1. Animal model and surgery
under an environmental scanning electron microscope (ESEM, FEI Based on the in vitro performances of Mg-xGe alloys, as-rolled
Quanta-200F, the Netherlands), operating in backscattering Mg-3Ge with the best mechanical properties and superior corro-
electron mode. Surface chemistry of the corroded samples was sion resistance was chosen to perform in vivo test. Twelve female
determined by using an imaging X-ray photoelectron spectrometer New Zealand rabbits (2–2.5 kg) were provided by Laboratory Ani-
(XPS, Axis Ultra, Kratos Analytical Ltd.) with Al Ka radiation. High mal Center of Peking University People’s Hospital (animal use per-
resolution narrow scanning was performed to determine the mit No.: SYXK (jing) 2011-0010). Animals were kept in separated
binding states of Mg 2p, O 1s, Ca 2p, P 2p, C 1s, and Na 1s. Fourier cages and health condition was verified by preoperative observa-
transform infrared spectroscopy (FTIR, Nicolet iS 50, Thermo tion 1 week before operation. Three animals were set as normal
Scientific) was utilized to identify the functional groups in the control and they were sacrificed at 3 months. The remaining nine
corrosion products. The spectrum was recorded from 4000 to animals all received implantation. Briefly, animals were anaes-
500 cm 1. Constituent phases in the corrosion product layer were thetized with 30 mg/kg pentobarbital sodium by ear vein injection.
determined by using an X-ray diffractometer (XRD, X’Pert Pro Operation sites beneath the bilateral knee-joint were shaved and
MPD). The test was performed by using a Mono-Capillary PreFix sterilized followed by decortication. A pre-drilled hole, 2.2 mm in
Module operated at 45 kV and 40 mA. diameter, was made on lower edge of the lateral tibial plateau by
using an orthopaedic drill. An as-rolled Mg-3Ge pin was inserted
2.6. Cytotoxicity into the predrilled hole. Then, 4  104 U penicillin was dropped
in the incision site before suturing. 40  104 U penicillin was
Cytotoxicity of the experimental materials was evaluated by intramuscular injected in case of postoperative infection and this
indirect cell assay, in which human osteoblast-like MG63 cell line injection lasted for 3 days. The anesthetic, surgical and
was used. Cells were cultured in minimum essential medium post-operative care protocols were examined by and fulfilled the
(MEM) supplied with 10% fetal bovine serum (FBS), 100 U/mL peni- requirements of the Ethics Committee of Peking University
cillin and 100 lg/mg streptomycin at 37 °C in a humidified atmo- People’s Hospital.
sphere with 5% CO2. Alloy extracts were prepared by using a
serum-free MEM with an extraction ratio of 1 cm2/mL under stan- 2.8.2. Postoperative observation, micro-CT, and histological evaluation
dard cell culture condition. After incubation for 72 h, the super- All animals were clinically examined for general condition and
natant fluid was withdrawn, centrifuged and stored at 4 °C in particular for significant signs of lameness, infection, subcuta-
before use. Ion concentrations and pH values of the extracts were neous emphysema formation and loss of appetite. Three animals
measured by inductively coupled plasma atomic emission spec- were sacrificed by over-dosage anesthesia at each time point
troscopy (ICP-AES, Leeman) and a pH meter (PB-10, sartorius), post-operation (1, 2 and 3 months), respectively. Rabbit tibias were
respectively. retrieved and the implantation section (1.5 cm) was truncated for
Methylthiazol tetrazolium (MTT) assay was adopted to evaluate micro-CT analysis and histological analysis. Bones were scanned in
the cytotoxicity of the alloy extracts according to ISO 10993- a self-built micro-CT device at a spatial resolution of 35 lm. Three-
5:2009(E) [41]. In brief, cells were seeded onto 96 well culture dimensional (3D) reconstruction was performed in Amira software
plate at a density of 3  103 cells per 100 lL. After incubation for (Amira 5.4.1, Visage Imaging). Retrieved bones were fixed in 10%
24 h to allow attachment, the medium was replaced by alloy formalin and dehydrated in gradient ethanol/distilled water mix-
extracts with normal MEM as control. After 1, 3 and 5 days’ incuba- tures. Then, they were embedded in methacrylate and sectioned
tion, 10 lL MTT was added to each well and another 4 h was kept. into 150 lm slices on a EXAKT system (EXAKT Apparatebau,
Afterwards, 100 lL formazan solubilizing solution (10% SDS in Norderstedt, Germany). The hard tissue sections were further
0.01 M HCl) was added into each well and left overnight in the grounded to 50–80 lm before staining with hematoxylin-eosin
incubator. Then, spectrophotometric absorbance of each well was (H&E) or toluidine blue (Guge Biological Technology Co., Ltd.).
measured by using a microplate reader (Bio-Rad 680) at 570 nm Details about the hard tissue slicing process could be found in
with a reference wavelength of 630 nm. our previous work [42]. Animal liver and kidney were also
retrieved for pathological examination by conventional H&E
2.7. Alkaline phosphate activity measurement staining.

Alkaline phosphate activity (ALP) of MG63 cells was determined


by hydrolysis of p-nitrophenyl phosphate (p-Npp) using a test kit 2.9. Statistical analysis
(Nanjing Jiancheng Bioengineering Institute, China). Briefly, after
5-day cultivation with extracts, cells were rinsed with a phosphate Data in this work were expressed as means ± standard
buffer solution (PBS, pH 7.4) for three times and then lysed in 0.1% deviation. Statistical analysis was performed with SPSS 18.0
Triton X-100 by freezing and thawing for 60 min at 37 °C. 30 lL of software (SPSS Inc., Chicago, USA). Differences between groups
the resulting cell lysates was then transferred to each well of a new were analyzed using one-way analysis of variance (ANOVA)
96 well plate, cultured with 50 lL carbonated buffer solution (pH followed by Tukey test. The statistical significance was defined as
= 10) and 50 lL substrate solution (4-amino-antipyrine) for 15 min d
p < 0.05.
424 D. Bian et al. / Acta Biomaterialia 64 (2017) 421–436

3. Results refined compared to Mg-1.5Ge, as shown in Fig. 1(b). Microstruc-


tures of the as-rolled Mg-Ge alloys were much more uniform com-
3.1. Microstructure pared to their as-cast counterparts. Area fractions of the eutectic
phases in the as-rolled Mg-Ge alloys (16.2 ± 1.1%, 25.1 ± 2.3% and
Fig. 1 shows the optical microstructures, average grain size and 35.9 ± 4.3%) were similar to their as-cast counterparts. Besides,
corresponding XRD patterns of experimental materials. All of the Eutectic phases were significantly refined after rolling.
alloys were composed of two phases, a-Mg and intermetallic com-
pound Mg2Ge with face centered cubic crystal system, as shown in 3.2. Mechanical property
Fig. 1(c). As-cast Mg-Ge alloys consisted of dendrites of a-Mg
matrix (bright white dendrites) and eutectic phases (a-Mg + Mg2 Fig. 2 displays the tensile mechanical behaviors of as-cast and
Ge, brown rod-like/network shaped) at the interdendritic regions, as-rolled Mg-Ge alloys. Tensile yield strengths (YS) and ultimate
as shown in Fig. 1(a). Area fraction of the eutectic phases increased tensile strengths (UTS) of the as-cast Mg-Ge alloys were in the
from 17.0 ± 1.3% in Mg-1.5Ge to 24.3 ± 2.9% in Mg-2.5Ge and then range of 36.6–51.4 MPa and 116.4–143.2 MPa, respectively.
to 33.9 ± 2.5% of Mg-3Ge. Meanwhile, average size of the eutectic As-cast Mg-2.5Ge and Mg-3Ge exhibited higher strengths and
phase continuously increased with increasing Ge addition. The elongations than as-cast Mg-1.5Ge. After rolling, YSs and UTSs of
a-Mg dendrites in Mg-2.5Ge and Mg-3Ge were significantly Mg-Ge alloys were significantly improved, however, elongations

Fig. 1. (a) Microstructures of as-cast and as-rolled Mg-Ge alloys; (b) average size of the a-Mg matrix and eutectic phase, d
p < 0.05 and * p < 0.01; (c) XRD patterns of
experimental alloys.
D. Bian et al. / Acta Biomaterialia 64 (2017) 421–436 425

Fig. 2. (a) Tensile stress-strain curves of experimental Mg-Ge alloys; (b) tensile yield strengths, ultimate tensile strengths and elongations of Mg-Ge alloys.

of Mg-1.5Ge and Mg-2.5Ge were impaired. Significantly improved For potentiodynamic polarization curves, the cathodic polarization
YS, UTS and elongation were found in as-rolled Mg-3Ge as they current reflected the severity of hydrogen evolution reaction on
were improved to 175.1 ± 5.1 MPa, 236.4 ± 0.7 MPa and 17.7 ± 2.5 platinum electrode while the anodic branch represented dissolu-
%, respectively. tion of magnesium. As-cast Mg-2.5Ge, as-cast Mg-3Ge and as-
rolled Mg-3Ge exhibited lower cathodic current density compared
3.3. Corrosion behavior to the remaining alloys. A little potential increase above the Ecorr
would cause a soaring increase in anodic current of Mg-1.5Ge.
OCP, EIS and Potentiodynamic polarization results were dis- The increase on the current of the anodic branches was reduced
played in Fig. 3(a)–(c) and corresponding electrochemical parame- for the samples with higher Ge content (2.5 and 3 wt%) when
ters were listed in Table 1. OCP values of all the experimental applying the same voltage disturbance above the Ecorr, implying
materials could generally stabilize after immersion for 1000 s. inhibited anodic reaction. The highest corrosion current density
OCP curves of as-cast Mg-1.5Ge and as-rolled Mg-1.5Ge fluctuated was found in as-cast Mg-1.5Ge and it was more than 2 orders of
frequently compared to the remaining ones, implying that it was magnitude of the lowest one, as-rolled Mg-3Ge. Among as-cast
difficult to reach a steady state for the formation and dissolution Mg-Ge alloys, corrosion current density significantly decreased
of the corrosion product film at the electrode/solution interface. with Ge content reaching 2.5 wt%. After hot rolling, reduced

Fig. 3. (a) OCP curves of Mg-Ge alloys in Hank’s solution; (b) potentiodynamic polarization curves of Mg-Ge alloys in Hank’s solution; (c) Nyquist plots of Mg-Ge alloys; (d)
volume of evolved hydrogen during immersion; (e) magnesium ion released into Hank’s solution; (f) Ge release after immersion.
426 D. Bian et al. / Acta Biomaterialia 64 (2017) 421–436

Table 1
Corrosion rates of Mg-Ge alloys in Hank’s solution.

Alloy Ecorr (V) icorr (lA/cm2) vcorr (mm/y) PH (mm/y)


as-cast Mg-1.5Ge 1.456 188.2 4.32 5.06
as-cast Mg-2.5Ge 1.603 1.2 0.03 0.16
as-cast Mg-3.0Ge 1.532 0.9 0.02 0.22
as-rolled Mg-1.5Ge 1.532 35.6 0.82 1.25
as-rolled Mg-2.5Ge 1.542 13.6 0.31 0.58
as-rolled Mg-3.0Ge 1.536 0.7 0.02 0.13

Ecorr: corrosion potential; icorr: corrosion current density; vcorr: corrosion rate calculated from polarization plots by using Tafel region extrapolation; PH: corrosion rate
calculated from hydrogen evolution.

Fig. 4. Surface morphologies (a) and cross section morphologies (b) of Mg-Ge alloys after immersion in Hank’s solution for 3 and 20 days. CP in (b) means corrosion products.

corrosion current densities were revealed for Mg-1.5Ge, while exhibited less dissolved Mg and Ge ions than their as-cast counter-
intensified corrosion was happened in Mg-2.5Ge. As-cast Mg-3Ge parts at day 10 and 20. Results of ion release were well consistent
and as-rolled Mg-3Ge owned the largest loop in the Nyquist plots, with the hydrogen evolution behaviors and all the corrosion tests
demonstrating their superior corrosion resistance among the demonstrated that as-rolled Mg-3Ge exhibited the best corrosion
alloys. Meanwhile, as-cast Mg-1.5Ge exhibited the smallest loop, resistance.
indicating its poor corrosion resistance. This finding was in agree- Fig. 4(a) shows typical corrosion morphologies of Mg-Ge alloys
ment with the potentiodynamic polarization results. after immersion in Hank’s solution for different durations. A corro-
Evolved hydrogen during immersion test was revealed in Fig. 3 sion product layer was formed on sample surface with some white
(d). At the initial stage of immersion (0–50 h), no much difference clusters/particles deposition. Some microcracks emerged on the
was found among the materials, all showing a low hydrogen evo- sample surface as a result of the shrinkage of corrosion product
lution rate. Thereafter, as-cast Mg-1.5Ge kept being corroded at a layer during dehydration. Severely localized corrosion was
relatively high corrosion rate level throughout the whole immer- observed on as-cast Mg-1.5Ge, which was consistent with its
sion period. However, other alloys maintained a low hydrogen evo- highest corrosion rate among all the alloys. Except for as-cast
lution trend and hydrogen volume gradually stabilized. Except for Mg-1.5Ge, corrosion product layers on the remaining alloys were
as-cast and as-rolled Mg-1.5Ge, volume of generated hydrogen was relatively uniform and compact. Judging from the cross-section
all well below 10 mL/cm2 after 20 days. As-cast Mg-2.5Ge, as-cast morphology in Fig. 4(b), differences in the corrosion behaviors
Mg-3Ge and as-rolled Mg-3Ge exhibited the slowest hydrogen could be easy to tell. As-cast Mg-1.5Ge and as-rolled Mg-1.5Ge
evolution rate, suggesting their superior corrosion resistance. both suffered severely localized corrosion. A thick corrosion pro-
Fig. 3(e) and (f) present released ion concentrations of Mg and duct layer with holes and cracks was observed above those deep
Ge in Hank’s solution at different immersion durations. Continuous corrosion sites. With increasing Ge addition, localized corrosion
dissolution of Mg and alloying element Ge into Hank’s solution was was suppressed as the boundary of the Mg-Ge matrix became more
found as their concentrations increased persistently with pro- regular and smoother. Thickness of the corrosion product layer
longed immersion time. Mg and Ge concentrations in Hank’s solu- decreased with the increase of added Ge, suggesting improvement
tion decreased with increased Ge content in the alloys, suggesting in the corrosion resistance. The thin corrosion product layer on
the positive effect of Ge on corrosion resistance. Hot rolling could as-cast and as-rolled Mg-3Ge could be hardly distinguished from
further improve corrosion resistance as as-rolled Mg-Ge alloys the Mg matrix and surrounding resin. This layer was closely
D. Bian et al. / Acta Biomaterialia 64 (2017) 421–436 427

Fig. 5. (a) FTIR spectra of as-rolled Mg-Ge alloys after immersion in Hank’s solution for 20 days; (b) XRD patterns after 20-d-immersion; (c) XPS analysis of the corrosion
products; (d) typical high resolution narrow scan results of Mg 2p, O 1s, Ca 2p, P 2p, C 1s and Na 1s among the corroded Mg-Ge alloys.

adherent to the Mg-3Ge matrix and could possibly provide protec- [46]. P showed a single peak at 133.4 eV. This peak was related
tion to the matrix. to PO34 or HPO24 , as also proved in FTIR. Generally, corrosion prod-
Fig. 5 shows the XPS, FTIR and XRD results of the corrosion ucts of Mg-Ge alloys were mainly composed of Mg(OH)2 accompa-
products on Mg-Ge alloys. Since the results of as-cast Mg-Ge alloys nied with small amount of CaCO3. Besides, a small quantity of
were similar to their as-rolled counterparts, so only the as-rolled carbonates, phosphates or their complex compounds might also
results were listed here. XPS results revealed that the corrosion exist.
products were mainly composed of Mg, O, Ca, P, C and limited
Na, as shown in Fig. 5(c). Ge was not detected mainly because of 3.4. Cytotoxicity and ALP activity
its minor content in this layer. FTIR indicated the presence of
H2O, OH , CO23 , PO34 and possible HPO24 , as depicted in Fig. 5(a) Fig. 6(a) and (b) present pH values and ion concentrations of the
[43,44]. Further XRD analysis confirmed the existence of Mg extract media. Due to the buffer agent in MEM, no much difference
(OH)2 in the corrosion products. Apparent diffraction peaks of was found in the pH values among all the extracts. Distinctive ion
CaCO3 were only detected on the Mg-1.5Ge sample. Due to a thick releasing behaviors observed in Hank’s solution were not observed
corrosion product layer on Mg-1.5Ge, strong peaks from Mg(OH)2 in MEM. Magnesium concentrations in various extracts were no
could be detected with minor signals from the a-Mg matrix. High longer apparently different. On the contrary, they seemed to be
resolution narrow scan of XPS also confirmed the presence of close to each other. Magnesium concentrations fluctuated between
CaCO3 on Mg-2.5Ge and Mg-3Ge, since one Ca 2p3/2 peak assigned 132 lg/mL and 160 lg/mL, while Ge concentrations were in the
to 347.4 eV, accompanied by C 1s at 289.7 eV and O 1s at 531.6 eV, range of 1.8–3.8 lg/mL. Ge concentrations in as-rolled Mg-2.5Ge
was detected [45]. No obvious CaCO3 signals on the XRD patterns and as-rolled Mg-3Ge extracts were apparently higher than the
of Mg-2.5Ge and Mg-3Ge should be ascribed to their limited remaining extracts.
amounts. Single Mg 2p peak at 50.4 eV along with O 1s at 532.9 Fig. 6(c) and (d) show the MG63 cell viability cultured in Mg-Ge
eV and C 1s at 289.7, as well as the presence of CO23 proved by alloy extracts and ALP activity after culturing for 5 days. According
FTIR, indicated possible presence of MgCO3 or hydrated MgCO3 to ISO 10993-5, cytotoxicity of biomaterials should be Grade 0 or
428 D. Bian et al. / Acta Biomaterialia 64 (2017) 421–436

Fig. 6. (a) pH values and (b) ion concentrations of the extract media; (c) MG63 cell viability and (d) ALP activity to Mg-Ge extracts.

Grade 1, which means cell viability should exceed 80% [41]. Except through micro-CT assessment, as depicted in Fig. 7. Generally,
for as-rolled Mg-3Ge, the remaining alloys all satisfied this crite- implantation of Mg-Ge pin did not cause observable bone damage
rion, exhibiting good cytocompatibility. Cell viability in as-rolled and normal bone morphology was observed, as revealed from the
Mg-3Ge extract was significantly lower than other extracts during 3D reconstruction (Fig. 7(a) and (b)).
the whole culture period. After 3 days, cell viability of as-rolled Continuous degradation of the implant could be observed and
Mg-3Ge was already lower than 80%, and further impaired viability no obvious signs of localized severe corrosion was found. In the
(47.9 ± 3.6%) was found at day 5. Cell viability of as-rolled Mg-Ge first two months, a few gas bubbles in quite limited size could be
alloys was lower than the as-cast ones, mainly due to the higher detected in medullary cavity of rabbit tibia, as marked in green
Mg and Ge concentrations in their extracts. Compared to control, color in Fig. 7(b). Up to 3 months, the previously existed gas bub-
increased ALP activity was observed for MG63 cells treated with bles around the implant could be no longer observed. During the
Mg-Ge extracts. As-cast Mg-1.5Ge owned the highest ALP concen- whole implantation period, Mg-Ge implant maintained intact, as
tration among all the experimental materials. Among the Mg-Ge shown in Fig. 8 (a). Surface morphology of the Mg-Ge implant
alloys in as-rolled condition, as-rolled Mg-3Ge exhibited the high- became rougher as time went on. Some mild and shallow corrosion
est ALP concentration. pits could be observed on implant surface after 3 months, implying
the appearance of minorly localized corrosion. Residual implant
3.5. In vivo performance volume calculated from 3D reconstruction is presented in Fig. 8
(b). At the initial stage after implantation (1 mon), the implant
Animals received implantation all survived and no obvious suffered a fast degradation as implant volume significantly
signs of lameness and loss of appetite were observed. No infection reduced with a volume loss of 15.4 ± 5.9%. Mitigated degradation
was found on wounds or peri-implant tissues through autopsy and was observed in the subsequent 2 months and the implant could
micro-CT examination. maintain 75.6 ± 3.2% of its original volume after 3 months.
Bone conductive property of Mg-Ge implant was also proved in
3.5.1. Micro-CT analysis the micro-CT assessment as newly formed bones could directly
In vivo degradation of Mg-Ge implants and osseointegration attach to the material surface and stretched along implant surface,
between implants and surrounding bones were characterized as clearly revealed in Fig. 7(c). Qualitatively, continuous
D. Bian et al. / Acta Biomaterialia 64 (2017) 421–436 429

Fig. 7. Micro-CT analysis of implant degradation and bone-implant interfaces. (a) 3D reconstruction of bone and implant; (b) 3D reconstruction with different parts rendered
with different colors, semitransparent gray for bone, orange for Mg-Ge implant, green for gas bubble; (c) 2D slice.

osseointegration and strengthened bonding between bone and tissue was presented between bone and implant surface. Enhanced
implant could be observed from the 2D and 3D reconstructions. osseointegration could be observed as directly contacted areas of
At 1 month, both ends of the implant were embed in the cortical bone and implant increased with prolonged implantation time,
bone, showing fast osseointegration. However, weak bonding as shown in Fig. 9(a) and (c). Bone-implant interfaces were closely
could be recognized as some voids could be found between related to the osseointegration and bonding strength. Typical fea-
implant and bone (not completely integrated). In the following tures at bone-implant interfaces were presented in Fig. 9(b). In
month (2 months), two end faces of the implant were gradually area A, implant was closely integrated with surrounding bones,
covered by newly formed bones, but still not directly contacted. showing well osseointegration. In area B, a transitional zone
After 3 months, bone and implant were well integrated as the existed between the implant and new bone. Signs of new bone
implant was tightly sealed by cortical bones. In addition, cortical penetration to the implant surface could be observed in this zone,
thickness surrounded the implant was thicker than that far from marked with white arrows. In area C, localized corrosion happened.
the implant, as revealed in Fig. 7(c). Bone-implant contact ratio Some surface fragmentations of the implant materials could be
(BIC) can quantitatively reflect the osseointegration status. In this found. A white gap without tissue filling presented between
study, BIC was calculated as the proportion of bone-implant implant and bone tissue, showing delayed osseointegration.
directly contacted area to the total implant surface in Amira soft-
ware. Continuously increased BIC value suggested enhanced
4. Discussion
osseointegration and bone-implant bonding, as shown in Fig. 8(c).
4.1. Microstructure, mechanical property and corrosion behavior
3.5.2. Histological analysis
Fig. 9 shows tissue responses adjacent to the Mg-Ge implant According to the Mg-Ge binary phase diagram, there is no solu-
after 1, 2 and 3 months with normal tissue as control. Clearly, con- bility of Ge in Mg and Mg2Ge phase must be precipitated, as shown
tinuous degradation of the Mg-Ge implant could be observed since in Fig. 10(a) [34]. Phase diagram also shows a high stability of this
the implant surface became much rougher with time. Generally, Mg2Ge intermediate phase (melting temperature 1117.4 °C),
the implant went through a macroscopic uniform corrosion mode substantially higher than Mg17Al12 phase ((melting temperature
without severely localized corrosion. No abnormality was found 455 °C)) in Mg-Al alloys [34]. In our Mg-xGe alloys, Ge content
with bone tissues surrounding the implant, showing good histo- was below 3.36 wt% (eutectic point). So, they all belonged to the
compatibility. No microscopic differences were noted in the tissue hypoeutectic group. Mg2Ge phase presented in the form of
structures of liver and kidney compared to normal controls, Sup- a-Mg + Mg2Ge hypoeutectic, surrounding the dendrites of primary
plementary Fig. 1. a-Mg phase. Area/volume ratio of those hypoeutectic phases
New bone was laid down directly onto the implant surface and increased with the increasing addition of Ge. Besides, net-like mor-
grew along the surface without interposed soft tissue layer, imply- phology was formed when Ge content exceeded 2.5 wt%, as shown
ing direct bonding of bone and implant. No scar tissue or fibrous in Fig. 1(a) and in Fig. 10(b). The improved strengths of Mg-Ge
430 D. Bian et al. / Acta Biomaterialia 64 (2017) 421–436

Fig. 8. (a) 3D reconstruction of the implant; (b) quantitative analysis of implant volume; (c) bone-implant contact ratio.

alloys were mainly derived from the second phase strengthening of Since Ge has no solubility in Mg and Mg-Ge alloys owned the same
Mg2Ge. Effects of solid solution strengthening could be neglected constituent phases of a-Mg and Mg2Ge, differences among their
since there is no solubility of Ge in Mg. Morphology of the eutectic corrosion behaviors should mainly depend on the volume fraction
structures might also have impact on alloy strength as the net-like and morphology of the Mg2Ge phase. The high-volume fraction
second phases could act as a skeleton in the transmission and dis- and continuously net-like eutectic phase containing Mg2Ge should
tribution of stress, possibly showing similar roles of concrete iron be responsible for the high corrosion resistance of Mg-2.5Ge and
and stones in reinforced concrete structures in architecture. Mg-3Ge alloys. At the same time, low volume fraction and inter-
The matrix a phase in magnesium alloys is normally anodic to mittent distribution of the eutectic phase (Fig. 10(b)) was believed
the second phases and is usually preferentially corroded. Eutectic to be closely related to the poor corrosion resistance of Mg-1.5Ge.
phase in our Mg-Ge alloys played two opposite roles during corro-
sion: (1) acted as cathode to the a-Mg matrix which accelerated 4.2. Toxicity of Ge and Mg-Ge alloys
the corrosion via galvanic corrosion; (2) acted as a barrier which
provided protection to the Mg matrix and thus slowed down the The use of germanium compounds in dietary supplements to
corrosion. Those two aspects interacted with each other and they promote health or cure diseases has become increasingly popular
were closely related to the volume fraction and morphology of in 1970s, especially in Japan. Until 1982, no systemic toxicity of
the eutectic phase [35]. Overall corrosion could be enhanced or Ge in man was reported [30]. Thereafter, numerous reports showed
suppressed depending on which process was the dominated one. that such supplements present a potential hazard to humans at
When the eutectic phase particles were in a low area/volume frac- high dose [28]. However, till now, Ge-containing health care
tion, they could not form a net-like morphology and could not pro- products, commonly available in organic forms and in capsules
vide protection of the matrix, as shown in the cross-section or tablets, as well as in powdered form, are still available in health
microstructure of corroded Mg-1.5Ge in Fig. 10(b). On the contrary, food stores and over the Internet. Cases of Ge induced toxicity in
micro-galvanic corrosion between the a phase and eutectic phase human were closely related to the high dosage (>100 times higher
would promote the evolution of severely localized corrosion [47]. than the estimated daily intake) and long-term supplement. Con-
However, for a high-volume fraction, the eutectic phase was nearly tamination of Ge in the inorganic forms, such as GeO2, was also
continuous like a net over the a matrix. Corrosion of the a phase suspected to be related to those clinical cases [28].
was then quite easily obstructed by the continuous eutectic phase, Ge compounds are relatively less toxic compared to other met-
and so the corrosion was greatly retarded, as demonstrated in the alloids and metals. However relatively high dose of germanium
cross-section microstructure of corroded Mg-3Ge in Fig. 10(b). dioxide and other inorganic Ge compounds can cause severe
D. Bian et al. / Acta Biomaterialia 64 (2017) 421–436 431

Fig. 9. Typical histological staining of hard tissue sections. (a) H&E staining; (b) local magnifications of specific areas in (a); (c) toluidine blue staining.

Fig. 10. (a) Phase equilibrium diagram of binary Mg-Ge system; (b) typical microstructures (upper) of as-cast Mg-1.5Ge and Mg-3Ge alloys, corresponding eutectic phase
morphologies (middle) and corresponding cross-section microstructures (under) after immersion in Hank’s solution for 20 days.

poisoning including death. The median lethal dose (LD50) of GeO2 synthesized in Japan in 1967) and Ge-lac-cit (germanium–lac
for mice ranges from 2020 mg/kg to 6300 mg/kg, with correspond- tate–citrate) are in the range of 0.7–23 mg/kg/d, equivalent to
ing values for rats of 1620 mg/kg and 3700 mg/kg, respectively 42–1380 mg/d in an adult weighing 60 kg [28], much higher than
[30]. It has been reported that daily intake of Ge from food in the daily intake.
human ranges from 0.4-3.4 mg, of which 96% is absorbed. At this Except for as-rolled Mg-3Ge, the other experimental Mg-Ge
concentration level, Ge would not cause toxicity [48]. Absorbed Ge alloys exhibited no toxicity to MG63 cells in this study. From the
is preferentially excreted in the urine in both humans and animals perspective of ion toxicity, the coexistence of 132–160 lg/mL Mg
[48,49]. Based on human case reports, the lowest observed adverse and 1.8–2.4 lg/mL Ge would not induce toxicity to MG63 cells.
effect levels of inorganic Ge, GeO2, organic Ge-132 ((GeCH2CH2 According to the results of Feyerabend et al. [50], 15 mM (360
COOH)2O3, a novel organogermanium compound originally lg/mL) Mg2+ would not cause cytotoxicity to MG63 cells. So,
432
Table 2
Some basic properties of typical essential/possibly essential elements in human body which can be used as alloying elements in magnesium.

Element Essential/trace Content Content in bone Blood Recommended Standard Physiological roles Perspective of material science Reference
in humana in human serum daily allowance electrode
level potential
(V) b
Mg Major, 21–28 g 60–65% in bones and teeth 0.75–1.2 320–400 mg 2.37 Activator of many enzymes; maintain muscle Poor mechanical properties of pure [63–65]
essential mM and nerve function; co-regulator of protein Mg hinder its use in load bearing
synthesis and muscle contraction; stabilizer applications; magnesium alloys with
of DNA and RNA sufficient mechanical properties are
needed
Ca Major, 1000 g 99% in bones and teeth 2.1–2.6 1000–1300 mg 2.87 Second messenger in signal transduction Improves strength of Mg (solid [66,67]
essential mM pathways; calcium ions as a cofactor in many solution strengthening, precipitation
enzymes; maintain proper bone formation hardening, grain refinement) and
creep resistance; reduces castability
Zn Trace, essential 2–4 g 29% in bones 12–16 12–15 mg 0.76 Zn has ubiquitous roles; participation in Improves strength, yet reduces [68,69]
mM metabolism of RNA and DNA, signal ductility in high concentrations (solid
transduction, and gene expression, also solution hardening, precipitation);
regulation in apoptosis; appearance in all improves castability

D. Bian et al. / Acta Biomaterialia 64 (2017) 421–436


enzyme classes
Mn Ultra-trace, 10–12 mg 43% in bones 10.9–78.3 1.8–5 mg 1.18 Necessary for development, metabolism, and Improves strength and ductility [70,71]
essential nM the antioxidant system; regulation of cellular (grain refinement); improves creep
energy, bone and connective tissue growth, resistance; improves corrosion
and blood clotting resistance in combination with
aluminium (precipitation that picks
up iron)
Fe Trace, essential 3–5 g 20 mg is transferred daily to 6.3–26.9 0.9 mg 0.44 The most important transition metal in all Fe is the most common impurity in [71–73]
the bone marrow for mM living organisms; iron-containing proteins are Mg; it is detrimental to the corrosion
erythropoiesis and is crucial for the transport and storage of oxygen resistance due to the formation of Fe-
efficiently recycled by containing particles
macrophages
Cu Ultra-trace, 84–126 0.53–5.69 ppm in bones 11–24 0.9–1.2 mg 0.34 Helps with the formation of collagen, Cu is a common impurity in Mg; a [74–76]
essential mg (1.4– mM increases the absorption of iron and plays a small amount of Cu has a beneficial
2.1 mg role in energy production; helps keep blood effect on the creep strength of
per kg of vessels, bones, nerves, and the immune magnesium die castings but strongly
body system healthy accelerates corrosion
weight)
Sr Trace, possibly 0.3 g 99.1% in bones 0.11–2.48 2 mg 2.89 No harmful effects of stable strontium in Improves strength and ductility [5,15,77,78]
essential mM humans; substitution for calcium in a small (grain refinement); improves creep
extent and deposition in bones; showing dose resistance and high temperature
dependent metabolic effect on bone; low strength
doses stimulated new bone formation
Si Trace, possibly 1–2 g No certain value. The highest 7.7–11.1 / 0.86 Involves in bone mineralization and Reduces ductility; improves creep [71,79,80]
essential levels are found in bone and mM prevention of osteoporosis, collagen resistance and high temperature
other connective tissues synthesis, and prevention of the aging of skin, strength (precipitation); reduces
such as, skin, nail, hair, overall condition of hair and nails corrosion resistance and castability
trachea, tendons and aorta
Sn Ultra-trace, 30 mg / 0.067– / 0.14 Tin-deficient diets in rat studies resulted in Improves strength, hardness and [5,21,81,82]
possibly 0.34 mM poor growth, reduced feeding efficiency, creep strength, however too much Sn
essential (in whole hearing loss, and bilateral (male) hair loss impairs strength and elongation; low
blood) concentration positively affects the
corrosion resistance
D. Bian et al. / Acta Biomaterialia 64 (2017) 421–436 433

higher Ge concentration (3.8 lg/mL) accompanied with higher

Elements listed below as ‘‘Essential in humans” are those listed by the (US) Food and Drug Administration as essential nutrients. Ultra-trace minerals are those elements with established, estimated, or suspected requirements
Mg concentration (148 lg/mL) in as-rolled Mg-3Ge extract

[27,33,85,86]
Reference

should be responsible for its low cell viability. Even though as-
[5,83,84]
rolled Mg-3Ge induced cytotoxicity to MG63 cells in vitro, the
in vivo tissue response and bone-implant osseointegration were
qualified. New bone could directly lay down onto the implant sur-
lattice structure); reduces corrosion
ductility/formability (change to bcc

face and grew along the surface, implying decent biocompatibility


high temperature strength (grain
Improves room temperature and
of this alloy. Owing to the capability of in vivo circulation system,
Reduces strength yet improves
Perspective of material science

local ion concentration surrounding the implant could be diluted.


refining and second phase

In order to minimize the huge gap between in vitro and in vivo


resistance and density

biocompatibility tests, 6–10 times dilution of extracts for


in vitro cytotoxicity test was recommended for biodegradable
magnesium-based materials [51]. In present study, local Mg and
strengthening)

Ge concentrations in vivo should be much lower than that


in vitro and that’s why as-rolled Mg-3Ge showed good biocom-
patibility in bone while exhibited cytotoxicity in vitro. During
the 3 months’ implantation, the total released Ge was only 0.58
mg, corresponding to 0.0064 mg per day under the assumption
Lithium is used in the treatment of bipolar

of a constant degradation rate. Daily Ge dissolution from the


including anticancer, anti-inflammation,
antioxidation, radioprotection and pro-

implant was far below its daily intake from normal diet and bio-
Organic Ge has been used as a dietary
supplement with therapeutic benefits

safety of this alloy could be guaranteed.

4.3. Mg-Ge alloys compared to other Mg-X (X: essential/possibly


essential elements in human health) alloy systems
Physiological roles

immune functions

Table 2 summarizes the basic properties of essential/possibly


essential elements in human body. As is well-known, elements
Ca and Mg are the major minerals in human body and the other
disorder

elements belong to the group of ‘‘trace elements” [52]. Ca seems


to be a proper alloying element in magnesium, especially for
orthopaedic implants, since it is the most abundant metal in
Standard electrode potential was excerpted from Electronic Science Tutor, G R Delpierre and BT Sewell, 1992–2012.
electrode
potential
Standard

human body and it is crucial to the health of bones and teeth.


3.04

But, poor formability and fast degradation might hinder the appli-
(V) b

0.24

cation of Mg-Ca alloys [7]. Strontium (Sr), zinc (Zn) and silicon (Si)
commonly found in bone tissues are believed to benefit for bone
daily allowance
Recommended

dietary intake

growth and health [53]. Strontium can replace calcium in the


0.4–3.4 mg/d
(estimated

of Ge)
/

(in plasma)
0.07 mM
serum

<1 mM
Blood

level

100 ng/g in bone


Content in bone

/
in human
Content

21–763
ppb

/
Essential/trace

Fig. 11. Comparison of the mechanical properties among various biodegradable


dietary trace
Ultra-trace,

Ultra-trace,

essential?)
in humana

Mg-X alloys (X: essential/possibly essential element in human health). Different


(possibly
essential
possibly

element

symbols represent different alloy systems and symbol color reflects the process
below 1 mg per day.
Table 2 (continued)

condition. Mechanical properties were deduced from the following references: Mg-
Ca alloys from Ref. [61,87,88], Mg-Zn alloys from Ref. [55,62,89,90], Mg-Sr alloys
from Ref. [15,16,18,91], Mg-Sn alloys from Ref. [20–23,82,92], Mg-Si alloys from
Element

Ref. [92,93], Mg-Mn alloys from Ref. [92], Mg-Li alloys from Ref. [24], Mg-Ge alloys
from Ref. [33] and from present study (purple marked). (For interpretation of the
Ge

references to colour in this figure legend, the reader is referred to the web version of
Li

b
a

this article.)
434 D. Bian et al. / Acta Biomaterialia 64 (2017) 421–436

Fig. 12. In vitro and in vivo corrosion rates of binary Mg-Ge alloys explored in this study compared to previously reported Mg-X alloys (X: essential/possibly essential element
in human health): (a) corrosion rates in Hank’s solution calculated from electrochemical tests; (b) corrosion rates in Hank’s solution calculated from immersion tests; (c)
corrosion rates in various bone tissues. In vitro corrosion rates: Mg-Ca alloys include as-cast Mg-1Ca [94]; Mg-Zn alloys include as-cast Mg-(1, 5, 7)Zn [55]; Mg-Sr alloys
include as-rolled Mg-(1, 2, 3, 4)Sr [15], and as-extruded Mg-(0.5, 1, 1.5, 2.5)Sr [18]; Mg-Sn alloys include as-cast Mg-(1, 3, 5, 7)Sn [20], as-extruded Mg-(1, 3, 5, 7)Sn [22], and
as-cast and as-rolled Mg-1Sn [92]; Mg-Si alloys include as-cast and as-rolled Mg-1Si [92]; Mg-Mn alloys include as-cast and as-rolled Mg-1Mn [92]; Mg-Li alloys include as-
extruded Mg-3.5Li and Mg-8.5Li [24]; Mg-Cu alloys include as-cast Mg-(0.05, 0.1, 0.25)Cu [13], and Mg-(0.03, 0.19, 0.57)Cu [14]; corrosion data of Mg-Ge alloys were results
in this study. In vivo corrosion rates of as-cast Mg-1Ca, as-extruded Mg-6Zn and as-rolled Mg-2Sr were derived from Ref. [61,62,15], respectively.

mineral of growing bones and thus lead to bone growth problems limited till now. Possibility of those alloys to be used as orthopae-
[54]. Thus, whether Mg-Sr alloys are applicable to children needs dic implants is still uncertain.
further research. Zn benefits for the strength and ductility simulta- Compared to other essential/possibly essential elements, bio-
neously. It is a proper alloying addition in both orthopaedic and logical information on Ge is quite limited. But, basically, it is a com-
cardiovascular applications [55,56]. Formability and corrosion mon trace element found in human diet and well controlled intake
resistance of Mg-Si alloys might be restricted by the brittle Mg2Si will not induce safety issues. Blood level of Ge is 3.99 lM (0.29 mg/
phase since Si has no solubility in Mg and Mg2Si must be precipi- L) in human serum, even higher than Li, Sn, Sr and Mn, as shown in
tated [56,57]. Fe and Cu are normally recognized as impurities in Table 2. Dietary intake of Ge is similar to some other nutrient ele-
magnesium since they are inevitably introduced from the raw ments. Standard electrode potential of Ge (Ge-Ge2+ + 2e, 0.24 V) is
materials or during the production process [58]. Mg-Fe and Mg- lower than Cu and higher than the remaining elements in Table 2,
Cu based alloys do not seem suitable for load bearing implants as for example, group neighbors Sn (Sn-Sn2+ + 2e, 0.14 V) and
they possess a fast degradation rate. However, they could be appli- Si (Si + 2H2O-SiO2 + 4H++4e, 0.86 V).
cable in special situations, such as Mg-Cu alloy in the treatment of An overall comparison of mechanical properties among various
osteomyelitis [13]. Among all the elements in Table 2, lithium (Li) Mg-X alloys (X represents for essential/possibly essential elements
is the only one who can change the hexagonal close-packed (hcp) in human health) is displayed in Fig. 11. Under the same processing
structure of Mg into body-centered cubic (bcc) type, thus improves condition, at as-cast or as-rolled state, mechanical properties of
ductility/formability. However, Li impairs strength. Mg-Li alloys to Mg-Ge binary alloys are comparable to other alloy systems.
be used as orthopaedic implants should be strengthened by other Strengths and elongations of as-cast Mg-Ge alloys in our study
alloying elements, such as Ca [59]. Generally, a common problem are higher than most of the remaining as-cast binary alloys, as
faced by Mg-Ca, Mg-Sr, Mg-Si and even Mg-Zn alloys is that they shown in Fig. 11. Among the limited binary alloys in as-rolled con-
exhibited fast degradation (>0.5 mm/y) either in vitro or in vivo dition, our as-rolled Mg-3Ge exhibits the highest strength and sig-
[15,60–62]. Information about Mg-Sn and Mg-Mn alloys is quite nificantly improved elongation. Generally, processing history has
D. Bian et al. / Acta Biomaterialia 64 (2017) 421–436 435

remarkable impacts on the mechanical properties and corrosion [7] Z. Li, X. Gu, S. Lou, Y. Zheng, The development of binary Mg-Ca alloys for use as
biodegradable materials within bone, Biomaterials 29 (10) (2008) 1329–1344.
behaviors. Performances of Mg-Ge alloys can be adjusted through
[8] M.H. Idris, H. Jafari, S.E. Harandi, M. Mirshahi, S. Koleyni, Characteristics of as-
various heat treatments and by deformation processing. In vitro cast and forged biodegradable Mg-Ca binary alloy immersed in Kokubo
corrosion rates of typical Mg-X alloys measured by electrochemical simulated body fluid, Adv. Mater. Res. 445 (2012) 301–306.
test and immersion test are displayed in Fig. 12 for comparison. [9] S. Zhang, X. Zhang, C. Zhao, J. Li, S. Yang, C. Xie, H. Tao, Z. Yan, Y. He, J. Yao,
Research on an Mg-Zn alloy as a degradable biomaterial, Acta Biomater. 6 (2)
Corrosion resistance of Mg-Ge alloys is quite outstanding. Among (2010) 626–640.
the Mg-X alloys, as-cast Mg-2.5Ge, as-cast Mg-3Ge and as-rolled [10] J. Li, Y. Song, S. Zhang, C. Zhao, F. Zhang, X. Zhang, L. Cao, Q. Fan, T. Tang, In vitro
Mg-3Ge exhibited the lowest corrosion rates (electrochemical), as responses of human bone marrow stromal cells to a fluoridated
hydroxyapatite coated biodegradable Mg–Zn alloy, Biomaterials 31 (22)
shown in Fig. 12(a). Superior corrosion resistance of those alloys (2010) 5782–5788.
was also verified in the immersion corrosion test, as shown in [11] E.M. Salleh, S. Ramakrishan, Z. Hussain, Characterization of binary Mg-Mn
Fig. 12(b). Besides, as-rolled Mg-3Ge exhibited much better alloy synthesized through mechanical alloying: effects of milling speed, Adv.
Mater. Res. 1087 (2015) 479–483.
in vivo corrosion resistance compared to typical Mg-Ca, Mg-Zn [12] G. Xie, H. Takada, H. Kanetaka, Development of high performance MgFe alloy
and Mg-Sr alloys of which orthopaedic implantation was as potential biodegradable materials, Mater. Sci. Eng. A 671 (2016) 48–53.
performed, as shown in Fig. 12(c). [13] Y. Li, L. Liu, P. Wan, Z. Zhai, Z. Mao, Z. Ouyang, D. Yu, Q. Sun, L. Tan, L. Ren,
Biodegradable Mg-Cu alloy implants with antibacterial activity for the
treatment of osteomyelitis: In vitro and in vivo evaluations, Biomaterials
5. Conclusions 106 (2016) 250–263.
[14] L. Chen, X. Fu, H. Pan, W. Peng, W. Lei, L. Tan, K. Wang, Z. Ying, Y. Ke, P.K. Chu,
Biodegradable Mg-Cu alloys with enhanced osteogenesis, angiogenesis, and
In this study, biodegradable Mg-based alloys with dietary trace long-lasting antibacterial effects, Sci. Rep. 6 (2016) 27374, https://doi.org/
element germanium were developed and their feasibility to be 10.1038/srep27374.
[15] X.N. Gu, X.H. Xie, N. Li, Y.F. Zheng, L. Qin, In vitro and in vivo studies on a Mg–
used as orthopedic applications was systematically investigated Sr binary alloy system developed as a new kind of biodegradable metal, Acta
through in vitro and in vivo evaluations. Mechanical properties of Biomater. 8 (6) (2012) 2360–2374.
Mg-Ge alloys were superior to many other biodegradable alloys [16] M. Bornapour, M. Celikin, M. Pekguleryuz, Thermal exposure effects on the
in vitro degradation and mechanical properties of Mg-Sr and Mg-Ca-Sr
strengthened with essential elements in human body. Besides, biodegradable implant alloys and the role of the microstructure, Mater. Sci.
they could be further improved by deformation processing, such Eng. C 46 (2015) 16–24.
as rolling and extrusion. Mg-Ge (>2.5 wt% Ge) alloys exhibited [17] H.S. Brar, J. Wong, M.V. Manuel, Investigation of the mechanical and degradation
properties of Mg–Sr and Mg–Zn–Sr alloys for use as potential biodegradable
excellent corrosion resistance since as-rolled Mg-3Ge exhibited a implant materials, J. Mech. Behav. Biomed. Mater. 7 (2012) 87–95.
slow degradation rate in vitro (vcorr = 0.02 mm/y, pH = 0.13 mm/y). [18] C. Zhao, F. Pan, L. Zhang, H. Pan, K. Song, A. Tang, Microstructure, mechanical
A proper degradation rate of 0.6 mm/y in vivo has also been veri- properties, bio-corrosion properties and cytotoxicity of as-extruded Mg-Sr
alloys, Mater. Sci. Eng. C 70 (Pt 2) (2017) 1081–1088.
fied. Quite limited H2 generated during implant degradation did
[19] A. Gil-Santos, I. Marco, N. Moelans, N. Hort, O.V.D. Biest, Microstructure and
not cause bone destruction and it could be completely absorbed degradation performance of biodegradable Mg-Si-Sr implant alloys, Mater. Sci.
in 3 months. Osseointegration of bone-implant was proved since Eng. C 71 (2017) 25–34.
new bone could directly lay down onto the implant and grew along [20] C. Zhao, F. Pan, S. Zhao, H. Pan, K. Song, A. Tang, Microstructure, corrosion
behavior and cytotoxicity of biodegradable Mg–Sn implant alloys prepared by
its surface. In conclusion, Mg-Ge alloys show great potential to be sub-rapid solidification, Mater. Sci. Eng. C 54 (2015) 245–251.
used as orthopaedic implants considering their good combinations [21] J. Kubásek, D. Vojtěch, J. Lipov, T. Ruml, Structure, mechanical properties,
of mechanical property, corrosion resistance, cytocompatibility corrosion behavior and cytotoxicity of biodegradable Mg–X (X=Sn, Ga, In)
alloys, Mater. Sci. Eng. C 33 (4) (2013) 2421–2432.
and histocompatibility. [22] C. Zhao, F. Pan, S. Zhao, H. Pan, K. Song, A. Tang, Preparation and
characterization of as-extruded Mg–Sn alloys for orthopedic applications,
Mater. Des. 70 (2015) 60–67.
Acknowledgements [23] J. Kubásek, D. Vojtěch, D. Dvorský, Structural and mechanical study on Mg–
xLM (x = 0–5 wt.%, LM = Sn, Ga) alloys, Int. J. Mater. Res. 107 (2016) 459–471.
This work was supported by the National Key Research and [24] W.R. Zhou, Y.F. Zheng, M.A. Leeflang, J. Zhou, Mechanical property,
biocorrosion and in vitro biocompatibility evaluations of Mg–Li–(Al)–(RE)
Development Program of China (Grant No. 2016YFC1102402), alloys for future cardiovascular stent application, Acta Biomater. 9 (10) (2013)
National Natural Science Foundation of China (Grant No. 8488–8498.
51431002), NSFC/RGC Joint Research Scheme (Grant No. [25] A.F. Cipriano, A. Sallee, R.-G. Guan, Z.-Y. Zhao, M. Tayoba, J. Sanchez, H. Liu,
Investigation of magnesium–zinc–calcium alloys and bone marrow derived
51361165101 and 5161101031) and NSFC-RFBR Cooperation
mesenchymal stem cell response in direct culture, Acta Biomater. 12 (2015)
Project (Grant No. 51611130054). 298–321.
[26] E. Zhang, L. Yang, J. Xu, H. Chen, Microstructure, mechanical properties and
bio-corrosion properties of Mg-Si(-Ca, Zn) alloy for biomedical application,
Appendix A. Supplementary data Acta Biomater. 6 (5) (2010) 1756–1762.
[27] L.G. Menchikov, M.A. Ignatenko, Biological activity of organogermanium
compounds (A review), Pharm. Chem. J. 46 (11) (2013) 635–638.
Supplementary data associated with this article can be found, in [28] S.H. Tao, P.M. Bolger, Hazard assessment of germanium supplements, Regul.
the online version, at https://doi.org/10.1016/j.actbio.2017.10.004. Toxicol. Pharm. 25 (3) (1997) 211–219.
[29] K. Yokoi, Germanium, Toxicity, Encyclopedia of Metalloproteins, Springer
Science+Business Media, New York, 2013.
References [30] A.G. Schauss, Nephrotoxicity and neurotoxicity in humans from
organogermanium compounds and germanium dioxide, Biol. Trace Elem.
[1] M.P. Staiger, A.M. Pietak, J. Huadmai, G. Dias, Magnesium and its alloys as Res. 29 (3) (1991) 267–280.
orthopedic biomaterials: a review, Biomaterials 27 (9) (2006) 1728–1734. [31] G.R. Belton, Y.K. Rao, Thermodynamic properties of Mg-Ge alloys, Metall.
[2] D. Zhao, F. Witte, F. Lu, J. Wang, J. Li, L. Qin, Current status on clinical Mater. Trans. B 2 (8) (1971).
applications of magnesium-based orthopaedic implants: a review from clinical [32] F. Islam, A.K. Thykadavil, M. Medraj, A computational thermodynamic model
translational perspective, Biomaterials 112 (2017) 287–302. of the Mg–Al–Ge system, J. Alloys Compd. 425 (1–2) (2006) 129–139.
[3] H. Hermawan, D. Dubé, D. Mantovani, Developments in metallic biodegradable [33] H.B. Liu, G.H. Qi, Y.T. Ma, X.G. Zhang, Microstructure evolution and mechanical
stents, Acta Biomater. 6 (5) (2010) 1693–1697. properties of Mg–Ge binary magnesium alloys, Mater. Res. Innov. 14 (2) (2013)
[4] A.M. Sammel, D. Chen, N. Jepson, New generation coronary stent technology–is 154–159.
the future biodegradable?, Heart Lung Circ 22 (7) (2013) 495–506. [34] J. Šerák, T. Kovalčík, D. Vojtěch, P. Novák, The influence of Ge on the properties
[5] Y.F. Zheng, X.N. Gu, F. Witte, Biodegradable metals, Mater. Sci. Eng. R. 77 of Mg alloys, Key. Eng. Mater. 647 (Suppl. 2) (2015) 72–78.
(2014) 1–34. [35] G. Song, A. Atrens, Understanding magnesium corrosion—a framework for
[6] J.W. Seong, W.J. Kim, Development of biodegradable Mg–Ca alloy sheets with improved alloy performance, Adv. Eng. Mater. 5 (12) (2003) 837–858.
enhanced strength and corrosion properties through the refinement and [36] ASTM-E8-04, Standard test methods for tension testing of metallic materials,
uniform dispersion of the Mg2Ca phase by high-ratio differential speed rolling, annual book of ASTM standards, American Society for Testing and Materials,
Acta Biomater. 11 (2015) 531–542. Philadelphia, PA, 2004.
436 D. Bian et al. / Acta Biomaterialia 64 (2017) 421–436

[37] H. Kuwahara, Y. Alabdullat, N. Mazaki, S. Tsutsumi, T. Aizawa, Precipitation of [64] R.R. Watson, V.R. Preedy, S. Zibadi, Magnesium in Human Health and Disease,
magnesium apatite on pure magnesium surface during immersing in Hank’s Humana Press, 2013.
solution, Mater. Trans. 42 (7) (2001) 1317–1321. [65] Y. Zheng, Magnesium Alloys as Degradable Biomaterials, CRC Press, 2015.
[38] ASTM-G102-89, Standard practice for calculation of corrosion rates and [66] M. Brini, D. Ottolini, T. Calì, E. Carafoli, Calcium in health and disease, Met. Ions
related information from electrochemical measurements, Annual book of Life Sci. 13 (4) (2013) 81–137.
ASTM standards Philadephia, American Society for Testing and Materials, [67] P. Govind, Role of calcium in human and animal health, Front. Cell. Infect.
Pennsylvania USA, 1999. Microbiol. Jigyasa 3 (2014) 39–45.
[39] ASTM G31-72, Standard practice for laboratory immersion corrosion testing of [68] G.P. Rink, L, Zinc and the immune system Zinc: immune system, Proc. Nutr.
metals, ASTM International, Philadelphia West Conshohocken, PA, 2004. Soc. 2000 (2000) 541–552.
[40] A.A. Ghoneim, A.M. Fekry, M.A. Ameer, Electrochemical behavior of [69] H. Tapiero, K.D. Tew, Trace elements in human physiology and pathology: zinc
magnesium alloys as biodegradable materials in Hank’s solution, and metallothioneins, Biomed. Pharmacother. 57 (9) (2003) 399–411.
Electrochim. Acta 55 (20) (2010) 6028–6035. [70] A.B. Santamaria, Manganese exposure, essentiality & toxicity, Indian J. Med.
[41] ISO 10993-5:2009(E), Biological evaluation of medical devices—Part 5: Tests Res. 128 (4) (2008) 484–500.
for in vitro cytotoxicity, International Organization for Standardization, 2009. [71] A. Sigel, H. Sigel, R.K.O. Sigel, Interrelations between Essential Metal Ions and
[42] P. Xiu, Z. Jia, J. Lv, C. Yin, Y. Cheng, K. Zhang, C. Song, H. Leng, Y. Zheng, H. Cai, Human Diseases, Metal Ions in Life Sciences, Springer, Dordrecht, 2013.
Tailored surface treatment of 3D printed porous Ti6Al4V by micro-arc [72] R. Crichton, Iron Metabolism: From Molecular Mechanisms to Clinical
oxidation for enhanced osseointegration via optimized bone in-growth Consequences, third ed., John Wiley & Sons, 2009.
patterns and interlocked bone/implant interface, ACS Appl. Mater. Int. 8 (28) [73] J.C. Dale, M.F. Burritt, A.R. Zinsmeister, Diurnal variation of serum iron, iron-
(2016) 17964–17975. binding capacity, transferrin saturation, and ferritin levels, Am. J. Clin. Pathol.
[43] P.K. Bowen, J. Drelich, J. Goldman, Magnesium in the murine artery: Probing 117 (5) (2002) 802–808.
the products of corrosion, Acta Biomater. 10 (3) (2014) 1475–1483. [74] S.S. Rajeswari, Role of copper in health and diseases, Int. J. Curr. Sci. 10 (2014)
[44] A.F. Anvari-Yazdi, K. Tahermanesh, S.M.M. Hadavi, T. Talaei-Khozani, M. E94–E107.
Razmkhah, S.M. Abed, M.S. Mohtasebi, Cytotoxicity assessment of adipose- [75] P. Trumbo, A.A. Yates, S. Schlicker, M. Poos, Dietary reference intakes: vitamin
derived mesenchymal stem cells on synthesized biodegradable Mg-Zn-Ca A, vitamin K, arsenic, boron, chromium, copper, iodine, iron, manganese,
alloys, Mater. Sci. Eng. C 69 (2016) 584–597. molybdenum, nickel, silicon, vanadium, and zinc, J. Am. Diet. Assoc. 101 (3)
[45] C.S. Gopinath, S.G. Hegde, A.V. Ramaswamy, S. Mahapatra, Photoemission (2001) 294–301.
studies of polymorphic CaCO3 materials, Mater. Res. Bull. 37 (7) (2002) 1323– [76] K. Jaouen, E. Herrscher, V. Balter, Copper and zinc isotope ratios in human bone
1332. and enamel, Am. J. Phys. Anthropol. 162 (3) (2017) 491–500.
[46] S. Ardizzone, C.L. Bianchi, M. Fadoni, B. Vercelli, Magnesium salts and oxide: an [77] S. Pors Nielsen, The biological role of strontium, Bone 35 (3) (2004) 583–588.
XPS overview, Appl. Surf. Sci. 119 (3) (1997) 253–259. [78] W.E. Cabrera, I. Schrooten, M.E. De Broe, P.C. D’Haese, Strontium and bone, J.
[47] J. Wang, Y. Jang, G. Wan, V. Giridharan, G.L. Song, Z. Xu, Y. Koo, P. Qi, J. Sankar, Bone Miner. Res. 14 (5) (1999) 661–668.
N. Huang, Flow-induced corrosion of absorbable magnesium alloy: in-situ and [79] E. Bissé, T. Epting, A. Beil, G. Lindinger, H. Lang, H. Wieland, Reference values
real-time electrochemical study, Corros. Sci. 104 (2016) 277–289. for serum silicon in adults, Anal. Biochem. 337 (1) (2005) 130–135.
[48] T.J. Chen, C.H. Lin, Germanium: environmental pollution and health effects, [80] R. Jugdaohsingh, Silicon and bone health, J. Nutr. Health Aging 11 (2) (2007)
Encyclopedia Environ. Health 16 (918) (2011) 927–933. 99–110.
[49] K. Yokoi, T. Kawaai, A. Konomi, Y. Uchida, Dermal absorption of inorganic [81] J. Chen, W.U. Shi-Hua, G.H. Guo, Y.M. Liu, Application of calcium nitrate in the
germanium in rats, Regul. Toxicol. Pharmacol. 52 (2) (2008) 169–173. determination of stannum in blood by graphite furnace atomic absorption
[50] F. Feyerabend, J. Fischer, J. Holtz, F. Witte, R. Willumeit, H. Drücker, C. Vogt, N. spectrometry, Chin. J. Health Lab. Technol. 26 (15) (2016) 2150–2153.
Hort, Evaluation of short-term effects of rare earth and other elements used in [82] H. Liu, Y. Chen, Y. Tang, S. Wei, G. Niu, The microstructure, tensile properties,
magnesium alloys on primary cells and cell lines, Acta Biomater. 6 (5) (2010) and creep behavior of as-cast Mg–(1–10)%Sn alloys, J. Alloys Compd. 440 (1–2)
1834–1842. (2007) 122–126.
[51] J. Wang, F. Witte, T. Xi, Y. Zheng, K. Yang, Y. Yang, D. Zhao, J. Meng, Y. Li, W. Li, [83] A.M. Pérez-Granados, M.P. Vaquero, Silicon, aluminium, arsenic and lithium:
K. Chan, L. Qin, Recommendation for modifying current cytotoxicity testing essentiality and human health implications, J. Nutr. Health Aging 6 (2) (2002)
standards for biodegradable magnesium-based materials, Acta Biomater. 21 154–162.
(2015) 237–249. [84] B. Sampson, Determination of low concentrations of lithium in biological
[52] C.D. Berdanier, D. Johanna, H. David, Handbook of Nutrition and Food, third samples using electrothermal atomic absorption spectrometry, J. Anal. At.
ed., CRC Press, 2013. Spectrom. 6 (6) (1991) 115–118.
[53] H.F. Li, X.H. Xie, Y.F. Zheng, Y. Cong, F.Y. Zhou, K.J. Qiu, X. Wang, S.H. Chen, L. [85] A. Shinohara, M. Chiba, Y. Inaba, Determination of germanium in human
Huang, L. Tian, L. Qin, Development of biodegradable Zn-1X binary alloys with specimens: comparative study of atomic absorption spectrometry and
nutrient alloying elements Mg, Ca and Sr, Sci. Rep. 5 (2015) 10719, https://doi. microwave-induced plasma mass spectrometry, J. Anal. Toxicol. 23 (7)
org/10.1038/srep10719. (1999) 625–631.
[54] Public Health Statement for Strontium, Agency for Toxic Substances & Disease [86] B.J. Kaplan, W.W. Parish, G.M. Andrus, J.S. Simpson, C.J. Field, Germane facts
Registry. about germanium sesquioxide: I. Chemistry and anticancer properties, J.
[55] S. Cai, T. Lei, N. Li, F. Feng, Effects of Zn on microstructure, mechanical Altern. Complement. Med. 10 (2) (2004) 337–344.
properties and corrosion behavior of Mg–Zn alloys, Mater. Sci. Eng. C 32 (8) [87] H. Du, Z. Wei, X. Liu, E. Zhang, Effects of Zn on the microstructure, mechanical
(2012) 2570–2577. property and bio-corrosion property of Mg–3Ca alloys for biomedical
[56] Y. Chen, Z. Xu, C. Smith, J. Sankar, Recent advances on the development of application, Mater. Chem. Phys. 125 (3) (2011) 568–575.
magnesium alloys for biodegradable implants, Acta Biomater. 10 (11) (2014) [88] H.R. Bakhsheshi-Rad, M.H. Idris, M.R. Abdul-Kadir, A. Ourdjini, M. Medraj, M.
4561–4573. Daroonparvar, E. Hamzah, Mechanical and bio-corrosion properties of
[57] M.E. Moussa, M.A. Waly, A.M. El-Sheikh, Effect of Ca addition on modification quaternary Mg–Ca–Mn–Zn alloys compared with binary Mg–Ca alloys,
of primary Mg2Si, hardness and wear behavior in Mg–Si hypereutectic alloys, Mater. Design 53 (2014) 283–292.
J. Magn. Alloys 2 (3) (2014) 230–238. [89] J. Kubásek, D. Vojtěch, Structural characteristics and corrosion behavior of
[58] D. Persaud, A. McGoran, Biodegradable magnesium alloys: a review of material biodegradable Mg–Zn, Mg–Zn–Gd alloys, J. Mater. Sci. Mater. Med. 24 (2013)
development and applications, J. Biomim. Biomater. Tissue Eng. 12 (1) (2011) 1615–1626.
25–39. [90] Q. Peng, X. Li, N. Ma, R. Liu, H. Zhang, Effects of backward extrusion on
[59] Y. Estrin, S.S. Nene, B.P. Kashyap, N. Prabhu, T. Al-Samman, New hot rolled Mg- mechanical and degradation properties of Mg–Zn biomaterial, J. Mech. Behav.
4Li-1Ca alloy: a potential candidate for automotive and biodegradable implant Biomed. Mater. 10 (2012) 128–137.
applications, Mater. Lett. 173 (2016) 252–256. [91] M. Bornapour, N. Muja, D. Shum-Tim, M. Cerruti, M. Pekguleryuz,
[60] M. Erinc, W. Sillekens, R. Mannens, R. Werkhoven, Applicability of existing Biocompatibility and biodegradability of Mg–Sr alloys: the formation of Sr-
magnesium alloys as biomedical implant materials, Magnesium Technology, substituted hydroxyapatite, Acta Biomater. 9 (2) (2013) 5319–5330.
The Materials Society Annual Meeting, 2009, pp. 209–214. [92] X. Gu, Y. Zheng, Y. Cheng, S. Zhong, T. Xi, In vitro corrosion and
[61] Z. Li, X. Gu, S. Lou, Y. Zheng, The development of binary Mg–Ca alloys for use as biocompatibility of binary magnesium alloys, Biomaterials 30 (4) (2009)
biodegradable materials within bone, Biomaterials 29 (10) (2008) 1329–1344. 484–498.
[62] S. Zhang, X. Zhang, C. Zhao, J. Li, Y. Song, C. Xie, H. Tao, Y. Zhang, Y. He, Y. Jiang, [93] E. Zhang, L. Yang, J. Xu, H. Chen, Microstructure, mechanical properties and
Research on an Mg–Zn alloy as a degradable biomaterial, Acta Biomater. 6 (2) bio-corrosion properties of Mg–Si(–Ca, Zn) alloy for biomedical application,
(2010) 626–640. Acta Biomater. 6 (5) (2010) 1756–1762.
[63] N.-E.L. Saris, E. Mervaala, H. Karppanen, J.A. Khawaja, A. Lewenstam, [94] B. Zhang, Y. Hou, X. Wang, Y. Wang, L. Geng, Mechanical properties,
Magnesium: an update on physiological, clinical and analytical aspects, Clin. degradation performance and cytotoxicity of Mg–Zn–Ca biomedical alloys
Chim. Acta 294 (1) (2000) 1–26. with different compositions, Mater. Sci. Eng. C 31 (8) (2011) 1667–1673.

You might also like