You are on page 1of 12

Materials Research Express

PAPER

Fabrication and properties of a biodegradable β-TCP/Zn–Mg bio-


composite
To cite this article: Lu Guoliang et al 2019 Mater. Res. Express 6 0865i1

View the article online for updates and enhancements.

This content was downloaded from IP address 35.176.47.6 on 21/08/2019 at 03:18


Mater. Res. Express 6 (2019) 0865i1 https://doi.org/10.1088/2053-1591/ab2861

PAPER

Fabrication and properties of a biodegradable β-TCP/Zn–Mg bio-


RECEIVED
10 April 2019
composite
REVISED
16 May 2019
ACCEPTED FOR PUBLICATION
Lu Guoliang1, Xu Guangquan2, Helian Bobo1, Liang Li1 and Liu Debao3
10 June 2019 1
School of Materials Science and Engineering, Tianjin University of Technology, Tianjin 300384, People’s Republic of China
2
PUBLISHED National Demonstration Center for Experimental Function Materials Education, Tianjin University of Technology, Tianjin 300384,
19 June 2019 People’s Republic of China
3
Tianjin Key Laboratory for Photoelectric Materials and Devices, Tianjin 300384, People’s Republic of China
E-mail: debaoliu@126.com

Keywords: β-TCP/Zn-1Mg composite, microstructure, mechanical properties, corrosion behavior, cell biocompatibility

Abstract
The good biocompatibility and moderate degradation rate of Zn-based alloys have attracted increasing
attention in biodegradable biomaterials field. However, the unsatisfactory mechanical properties of
pure zinc limited its biomedical applications. In this study, Zn-1Mg alloy and 1vol%β-TCP/Zn-1Mg
composite were fabricated using stirring casting assisted by ultrasonic vibration. Microstructure,
mechanical properties, corrosion behavior and osteotoxicity of Zn-1Mg alloy and 1vol%β-TCP/Zn-
1Mg composite were tested and /evaluated. In comparison with the matrix alloy, the addition of
1vol%β-TCP refined the grain size. The yield strength(YS), ultimate tensile strength(UTS) and
elongation of the as-extruded composite were 249.3 MPa, 330.1 MPa, and 11.7% respectively, which
increased by 10.2%, 9.9%, and 101.7% respectively. The corrosion rate of 1vol%β-TCP/Zn-1Mg
composite in SBF is 46 μm/a. Corrosion resistance of the composites was improved after ultrasonic
treatment. The proliferation rate of MC3T3-E1 cell for Zn-1Mg alloy and 1vol%β-TCP/Zn-1Mg
composite extracts were above 80% after 1–5days in culture. As the culture time increased, the 1vol%
β-TCP/Zn-1Mg composite showed superior cytocompatibility compare to Zn-1Mg alloy.

In recent years, the research on biodegradable metallic materials attracted increasing attention in biomedical
field, because this kind of materials can avoid secondary surgery and reduce the psychological and economic
burden on patients in addition to their good mechanical properties, workability and good biocompatibility
[1–3]. Currently, biodegradable metallic materials mainly can be categorized in the following three groups base
on major alloying element:iron-based, magnesium-based and zinc-based materials. Magnesium alloys which
have excellent biosafety and mechanical compatibility have been widely investigated and developed as
degradable metals. However, because of the actively chemical nature of magnesium and its alloys, their
degradation rates in corrosive environments are extremely fast, which can lead to the loss of the mechanical
properties when they are in service uses [4–6]. On the contrary, the corrosion rate of the Fe and its alloys is much
lower than clinical needs, which makes the Fe-based implants not suitable for utilization as biodegradable
implants in human body. Moreover, their own ferromagnetism also limited biomedical applications under
Magnetic Resonance Imaging(MRI) environment. On the other hand, the chemical activity of Zinc is lower than
that of magnesium but higher than iron. Compared with magnesium and iron, zinc and its alloys have more
suitable corrosion rate for biodegradable applications, which can significantly meet the clinical needs. Therefore,
Zinc has attracted increasing attention as promising candidates for biodegradable applications in recent years
[7, 8]. However, the tensile strength and elongation rate of as-cast pure zinc are only 120 MPa and 2%
respectively, which is insufficient to meet the mechanical requirement of implant materials. Hence,
improvement of the mechanical properties of zinc is essential for biodegradable applications. Magnesium has
been proved as an effective alloying element for Zinc to improve the strength and biocompatibility [9]. The bone
healing process may also be improved by the addition of magnesium due to the positive effect of magnesium on
bone growth [10, 11]. Therefore, we have selected Zn–Mg alloying system in our research. At the meantime, β-

© 2019 IOP Publishing Ltd


Mater. Res. Express 6 (2019) 0865i1 L Guoliang et al

Figure 1. The preparation schematic diagram of β-TCP/Zn-1Mg composites (a) Mechanical stirring, (b) The ultrasonic treatment.

tricalcium phosphate (β-TCP) which often used in bone biocomposites as a reinforced phase has excellent
biocompatibility, bioactivity and osteoconductive properties. Furthermore, compared with HAP that shows
very low solubility in body fluid, β-TCP has high dissolution rate therefore it is useful to achieve complete
degradation implant materials [12, 13].
In this study, Zn-1wt%Mg (as Zn-1Mg) alloy and 1vol%β-TCP/Zn-1wt%Mg (as β-TCP/Zn-1Mg)
composite were fabricated using stirring casting assisted by ultrasonic vibration followed by hot extrusion. the
microstructure, mechanical properties, corrosion behavior and cytocompatibility of as-cast and as-extruded
composites were tested and evaluated. These experimental results may provide direct guidelines regarding the β-
TCP/Zn-1Mg composite apply for biodegradable applications in the future.

1. Experimental procedures

1.1. Composite fabrication


In this study, pure zinc (99.995%), Zn–Mg intermediate alloy (20wt%Mg) and β-TCP (99%) particles of
0.5–15 μm in size were selected as raw materials. The preparation schematic diagram of 1 vol% β-TCP/Zn-1Mg
composites is shown in figure 1. Firstly, pure zinc and Zn–Mg intermediate alloy were melted at 500 °C under
the protective atmosphere of N2+SF6. Then the melt was mechanically stir in melted condition for 5 min
followed by 1 min standing. At the same time, the preheated β-TCP particles were added into the melt. After
mechanical stirring ultrasonic probe consisted of a transducer coupled with a maximum power of 1 KW and
frequency of 20 kHz was imposed to the melt for 5 min. At last, the melt was pour into a preheated round-shaped
steel mold to obtain a β-TCP/Zn-1Mg composite. The experimental procedure for the fabrication of Zn-1Mg
alloy is almost the same as that of β-TCP/Zn-1Mg composite except the absence of β-TCP. The mechanical
stirring condition and ultrasonic treatment condition is the same.
After stirring casting assisted by ultrasonic vibration, annealing at 350 °C was conducted for 24 h in order to
homogenize the cast structure followed by water quenching. Annealed samples were subsequently extruded at
250 °C with an extrusion ratio of 16:1.

1.2. Morphology and structural characterization


The samples were ground and polished following standard metallographic procedures and etched in a 5 wt%
HNO3 solution. The microstructure of Zn-1Mg alloys and β-TCP/Zn-1Mg composites were observed using an
OLYMPUS GX51 metallographic microscope and an environmental field emission scanning electron
microscope (FE-SEM, Quanta FEG 250) equipped with an Energy Dispersive Spectrometer (EDS). The phase
constitution was analyzed by x-ray diffraction (XRD, Rigaku D/max/2500PC, Japan), using Cu Kα radiation
with the scanning range 2θ from 10° to 80°, and the scanning speed of 5 °/min.

1.3. Mechanical property


Mechanical properties were evaluated by tensile tests. The tensile mechanical properties of β-TCP/Zn-1Mg
composites were measured by using an Instron DDL10 universal testing machine at a constant cross-head speed
of 0.5 mm min−1. The tensile tests were in accordance with ASTM: E8/E8M standards [14]. The tensile property
data were based on the average of 3 tests. The fracture morphology was observed by SEM.

2
Mater. Res. Express 6 (2019) 0865i1 L Guoliang et al

1.4. Corrosion behavior


1.4.1. In vitro immersion test
The specimens were polished with 3000 grit SiC paper before washing with ethanol for 5 min in ultrasonic bath
to remove surface particles, and then rinsed with distilled water for 5 min. The Zn and Zn-based extruded alloy
samples were weighed after drying and immersed for 336 h in simulated body fluid (SBF) The pH value of the
SBF was adjusted to 7.4, and the temperature was kept at 37 °C in a water bath. After different immersion
periods, the samples were removed from SBF and then carefully rinsed with distilled water before they are dried
by warm air. The corrosion products were then removed by a chromate solution for the surface morphology
observation by SEM.
The corrosion rate of the immersed specimens could be calculated according to equation (1):

k·W
Rc = (1)
A·t·D

Rc is the corrosion rate (mm/year), k is constant 8.76×104 , W is the weight loss (g), A is sample area (cm2),
t is immersion time (h), and D is sample density (g cm−3). Three parallel samples are set at each time, for
calculation and averaging.

1.4.2. Electrochemical corrosion test


The electrochemical test analysis is carried out in a simulated body fluid using a German Zahner-Zennium
electrochemical workstation. Zn-1Mg alloys and β-TCP/Zn-1Mg composites were cutted into size of Φ15
mm×2.5 mm and then polished with 3000 grit SiC paper, before washing with ethanol for 5 min followed by
dry to test the exposed area which is 1 cm2. Dynamic polarization curve (I/E) and AC impedance spectroscopy
(EIS) tests were conducted to evaluate the electrochemical corrosion behavior. All samples are immersed in
simulated body fluids for 30 min before test. The electrochemical impedance spectroscopy (EIS) experiments
were performed in a three-electrode cell, using a graphite electrode as the counter electrode and a saturated
calomel electrode (SCE) as the reference electrode with the frequency range of 100 kHz to 1 mHz and the
sinusoidal signal perturbation of 10 mV. The simulated body fluid is used as the solution medium for
electrochemical test. The potentiodynamic polarization test is performed at a scanning speed of 1 mV s−1 after
the EIS test.

1.5. Cytotoxicity test


Mouse fibroblasts (L-929 cells) and mouse osteoblasts (MC3T3-E1) were cultured in the RPMI-1640 medium
(Gibco, US) with 10% (volume fraction) fetal bovine serum (FBS), 100 U mL−1 penicillin and 100 mg/ml
streptomycin at 37 °C in a humidified atmosphere with 5% CO2. The cytotoxicity tests were carried out by
indirect contact. The extracts were prepared according to ISO 10993-5: 1999 [15]. The corrosion products were
sterilized by irradiation. The extraction medium was diluted with medium at a ratio of 1/2, 1/4, 1/8, 1/16,
respectively. The control group used 10% serum+90% RPMI-1640 medium. The cells were incubated in 96-
well flat-bottomed cell culture plates at 50 cell L−1 medium in each well and incubated for 24 h to allow
attachment. The medium was then replaced with 100 μl extracts. The cells was incubated in a humidified
atmosphere with 5% CO2 at 37 °C for 1, 3 and 5 d, respectively, and then the cell viability was measured by MTT
(methyl thiazolyl tetrazolium) method. Calculate the relative proliferation rate (RGR) of the cells according to
RGR%=(OD mean of the control group/OD mean of the negative control group)×100%. The RGR value is
converted to 0 to V based on the grading standard to assess the toxicity of materials on cells: 100% or more=0
grade, 75%∼99%=I, 50%∼74%=II, 25%∼49%=III, 1%∼24%=IV, 0=V.

2. Experimental results and discussion

2.1. Phase and microstructure


Figure 2 showed XRD pattern of the as-cast Zn-1Mg alloy and the β-TCP/Zn-1Mg composite. The result
revealed that the alloy and composite mainly consist of α-Zn and Mg2Zn11 phase. The peaks from MgZn2 phase
were also confirmed in as-cast Zn-1Mg alloy and the β-TCP/Zn-1Mg composite. Since α-Zn phase is
precipitated first in the cooling and solidification process, the precipitation of α-Zn phase may causes the
composition of liquid phase to deviate from the eutectic point, resulting in the conversion of remaining liquid
phase from eutectic to peritectic transformation during cooling process. Gong et al reported that as-cast Zn-1Mg
possessed both the eutectic phase that consist of α-Zn and Mg2Zn11 as well as pre-solidified the α-Zn phase [10].
The Zn-0.8Mg reported by Kubásek J et al reported similar phase constitution and microstructure in Zn-0.8Mg
alloy [11]. No remarkable β-TCP diffraction peaks were detected due to the small amount of β-TCP particles
added.

3
Mater. Res. Express 6 (2019) 0865i1 L Guoliang et al

Figure 2. XRD pattern of as-cast Zn-1Mg alloy and β-TCP/Zn-1Mg composite.

Figure 3. SEM photographs and EDS analysis of the as-cast microstructure of Zn-1Mg alloy and β-TCP/Zn-1Mg composite (a) Zn-
1Mg, (b) β-TCP/Zn-1Mg.

Figure 3 showed the SEM images and corresponding EDS analysis of the as-cast microstructure of Zn-1Mg
alloy and β-TCP/Zn-1Mg composite. The α-Zn matrix and eutectic phase was confirmed as-cast Zn-1Mg alloy
as shown in figure 3(a). EDS analysis revealed that the eutectic phase contained only Zn and Mg with the atomic
ratio of Zn/Mg is 14.22 which is quite different from that of Mg2Zn11. Therefore, we assumed that the eutectic
phase should consist of α-Zn and Mg2Zn11, which was consistent with the XRD results. Figure 3(b) showed SEM
image of the as-cast β-TCP/Zn-1Mg composite. Due to the amount of β-TCP was quite limited, the
microstructure was nearly the same as that of the as-cast Zn-1Mg alloy. EDS analysis confirmed a small amount
of β-TCP existed inside the matrix.

4
Mater. Res. Express 6 (2019) 0865i1 L Guoliang et al

Figure 4. Microstructure of Zn-1Mg alloy and β-TCP/Zn-1Mg composite: (a) as-cast Zn-1Mg, (b) as-extruded Zn-1Mg, (c) as-cast β-
TCP/Zn-1Mg, (d) as-extruded β-TCP/Zn-1Mg.

Figure 4 showed OM photographs of as-cast and as-extruded Zn-1Mg alloy as well as β-TCP/Zn-1Mg
composite. OM observation revealed that the Zn-1Mg alloy contains dendrites and also some equiaxed grain
containing small particles. As-cast Zn-1Mg consists of dendrites, primary grains and eutectic structure were
shown in figure 4(a). Figure 4(b) showed OM photograph of as-extruded Zn-1Mg alloy observed from the
direction perpendicular to the extrusion direction. We found extrusion process not only remarkably refines the
grains size of the Zn-1Mg alloy, but also causes the dissolution of Zn+Mg2Zn11 eutectic phase which
transformed to small size Mg2Zn11 precipitate distributed along the grain boundary. The large dendrites
observed in as-cast Zn-1Mg alloy are significantly reduced in β-TCP/Zn-1Mg composite. β-TCP particles
mainly distributed along the grain boundaries, in spite of a small amount of β-TCP particles dispersed
intragranularly. During solidification process, primary α-Zn grains which show the highest melting point in Zn-
1Mg system nucleated and grew firstly. Accompanying with the grain growth of α-Zn, Mg segregation was
occurred during solidification due to the limited solubility in α-Zn. Mg-rich area formed in the grain boundary.
At the meantime, the β-TCP particles were repelled to the last liquid phase during cooling process. When the
temperature dropped to the eutectic temperature, the eutectic reaction could occur. This mechanism resulted in
the eutectic phase and the β-TCP particles localized at the grain boundaries. In comparison with the matrix
alloy, the addition of 1vol% β-TCP refined the average grain size as shown in figures 4(a) and (c). Obvious grain
refinement was confirmed in as-extruded β-TCP/Zn-1Mg composite as shown in figure 4(d). During hot
extrusion, β-TCP particles give rise to high level of local lattice misorientation around the particle, which can act
as potent nucleation sites for new recrystallized grains in a process known as particle stimulated nucleation
(PSN) []. Moreover, β-TCP particles also can inherent the migration of the grain boundaries as the Orowan-type
strong obstacles, thereby contributing to grain refinement in β-TCP/Zn-1Mg composite.

2.2. Mechanical properties


Figure 5 showed stress-strain curve of the as-extruded Zn-1Mg alloy and β-TCP/Zn-1Mg composite. the
Ultimate tensile strength (UTS), yield strength (Y.S.) and elongation of the as-extruded Zn-1Mg alloy are
300.5 MPa, 226.2 MPa and 5.8%, respectively. Compare to Zn-1Mg alloy, 1vol% β-TCP/Zn-1Mg composite
showed far superior mechanical properties. The UTS, Y.S and elongation of β-TCP/Zn-1Mg composite increase

5
Mater. Res. Express 6 (2019) 0865i1 L Guoliang et al

Figure 5. Tensile stress and strain curve of as-extruded Zn-1Mg alloy and Zn-1Mg-β-TCP composite.

Figure 6. Fracture surfaces of the Zn-1Mg alloy and β-TCP/Zn-1Mg composite after tensile tests: (a) Zn-1Mg, (b) β-TCP/Zn-1Mg.

to 330.1 MPa, 249.3 MPa and 11.7% respectively. With respect to the microstructural analysis, the grain
boundary strengthening and precipitation strengthening due to β-TCP addition are presumed to be responsible
for the strength and ductility enhancement in β-TCP/Zn-1Mg composite. The addition of ultrafine β-TCP
particles can refine grains during solidification and subsequent hot extrusion process. Grain refinement not only
improves the strength of Zn matrix, but also the plastic deformation ability of the composition. The ultrafine
precipitates dispersed along the grain boundary were formed during extrusion process caused by dissolution of
eutectic phase and agglomerated β-TCP particles which can be observed in as-cast composite. These ultrafine β-
TCP precipitates lead to precipitate strengthening in β-TCP/Zn-1Mg composite. Therefore, it is reasonable to
assume that addition of β-TCP particles contributed to grain boundary strengthening as nucleation agent and
precipitation strengthening as orowan-type obstacle resulted in the improvement of strength and ductility in β-
TCP/Zn-1Mg composites. Figure 6 showed the SEM images of fracture surface for as-extruded Zn-1Mg alloy
and β-TCP/Zn-1Mg composite. Relatively brittle fracture mode composed of large area cleavage facets and
limited amount of small size dimples was confirmed by fracture surface observation of Zn-1Mg alloy. In β-TCP/
Zn-1Mg composites, relatively smaller size cleavage facets and obvious dimples were observed revealing the
improvement of ductility.

2.3. Corrosion behavior


2.3.1. Immersion test analysis
Figure 7 showed the corrosion rate of the as-extruded Zn-1Mg alloy and the β-TCP/Zn-1Mg composite as a
function of immersion time. It can be seen that the change trend of corrosion rate of the Zn-1Mg alloy and the β-
TCP/Zn-1Mg composite is the same throughout the immersion period, i.e. at the early stage of immersion
testing the corrosion rate was relatively high and decreased with increasing immersion time. Eventually, the
corrosion rate decreased from 3.15 mm/a for 1 day immersion to 0.046 mm/a for 30 days. It is suppose to be
that large area of α-Zn was exposed to the SBF at the early stage of immersion, resulting in the α-Zn was rapidly
corroded. As corrosion progresses, the corrosion products formed a protective layer interrupted the contact

6
Mater. Res. Express 6 (2019) 0865i1 L Guoliang et al

Figure 7. Corrosion rate of Zn-1Mg alloy and β-TCP/Zn-1Mg composite at different immersion time in SBF solution.

between α-Zn and corrosion environment(SBF) thereby reduced corrosion rate. The 1% β-TCP/Zn-1Mg
composite indicated the same trend to Zn-1Mg alloy in corrosion rates with different immersion time. In
addition, for 1–21days immersion, the corrosion rate of the β-TCP/Zn-1Mg composite was slightly higher than
that of the Zn-1Mg alloy. On the contrary, the corrosion rate of Zn-1Mg alloy and β-TCP/Zn-1Mg composite
were almost same after 30 days immersion. We assumed that β-TCP particles induced microstructure and
surface passivability variation caused different corrosion behavior between Zn-1Mg alloy and β-TCP/Zn-1Mg
composite. At the early stage of immersion, because the binding strength between agglomerated β-TCP particles
and α-Zn matrix was quite weak, α-Zn matrix around the agglomeration was easily corroded accompanying
with the detachment of agglomerated β-TCP resulting in an increased corrosion rate. On the other hand, β-TCP
particles could induce deposition of Ca2+ and P5+ ions in SBF, which promotes the formation of the surface
passivation layer contributed to the increase of corrosion resistant at later stage of immersion test(30days) .
Figure 8 showed the surface morphologies of Zn-1Mg alloy and β-TCP/Zn-1Mg composite samples after
immersion in SBF for 1 day and 30days. It can be seen that the surface morphology of both Zn-1Mg alloy and β-
TCP/Zn-1Mg composite after 1 day immersion does not change significantly. On the contrary, After a
prolonged immersion period (30 days), the surfaces of all specimens were covered by a layer of corrosion
products as shown in figures 8(c), (d). SEM observation revealed that the needle-like corrosion products were
deposited on the corrosion surface of the β-TCP/Zn-1Mg composite (as shown in the enlarged inset in
figure 8(d)). The corrosion behavior of the β-TCP/Zn-1Mg composite in SBF could be summarized as the
following corrosion reaction equation:
Zn2 + + 2H2 O = Zn (OH)2 + H2 (Total recation) (2)
10Ca2 + + 6PO34- + 2OH-  2Ca10(PO4)6 (OH)2 (3)
Therefore, the main composition of the corrosion product layer on the composite surface are Zn(OH)2 and
Ca10(PO4)6(OH)2, which can prevent the further corrosion from corrosion environment (SBF) and thus
reduced corrosion rate. However, Due to the Cl− contained in the simulated body fluid, however, the insoluble
corrosion product can react with Cl− which was contained in SBF to form soluble product, thereby destroying
the integrity of the corrosion layer and causing the detachment of the corrosion product layer from the
composite surface.

2.3.2. Electrochemical corrosion analysis


Figure 9 showed the potentiodynamic polarization curves of Zn-1Mg alloy and β-TCP/Zn-1Mg composite in
SBF. These polarization curves are composed of cathode branch and anode branch, and cathode branch is
mainly related to the hydrogen-evolution reaction on the surface of the β-TCP/Zn-1Mg composite electrode,
while the anode branch is related to the anodic dissolution of the β-TCP/Zn-1Mg composite electrode. As can
be seen in figure 9, both the Zn-1Mg alloy and the β-TCP/Zn-1Mg composite were obviously passivated in SBF.
Passivation is a process in which a metal surface was converted from an active state to an inactive state by
forming a protective layer. In SBF solution, passivation of Zn-based material usually could divide into three
steps: firstly, zinc is oxidized to form Zn(OH)−3 or Zn (OH)- 2
4 ; secondly, they are dissolved in the electrolyte to
−3
product Zn(OH) or Zn(OH); finally, when the electrolyte can no longer dissolve ZnO or Zn(OH)2, the

7
Mater. Res. Express 6 (2019) 0865i1 L Guoliang et al

Figure 8. The surfaces of the Zn-1Mg alloy and β-TCP/Zn-1Mg composite exposed to the SBF with different immersion time (a) Zn-
1Mg alloy 1 day, (b) β-TCP/Zn-1Mg composite 1 day, (c) Zn-1Mg alloy 30 day, (d) β-TCP/Zn-1Mg composite 30 day.

Figure 9. The potentiodynamic polarization curves of Zn-1Mg alloy and β-TCP/Zn-1Mg composite soaked in the SBF.

protective solid film formed and passivated the electrode surface. The potentiodynamic polarization process of
Zn-1Mg alloy and β-TCP/Zn-1Mg composite in SBF is controlled by activation in a large voltage range, and
exhibits different potentiostatic polarization behavior. The corrosion potential was negatively shifted after the
addition of β-TCP to Zn-1Mg matrix. As we all known, higher corrosion potential means better corrosion
resistance. Hence, the potentiodynamic polarization results reveal that β-TCP can accelerate corrosion in Zn-
1Mg alloying system. That is due to the addition of β-TCP particles as reinforcement introduce the ceramic
reinforcement/metallic matrix interface which showed weak interface bonding. The interface has a detrimental
effect on the integrity and continuity of the passivation film on the composite surface, resulting in higher
corrosion rate of the β-TCP/Zn-1Mg composite compare to the Zn-1Mg alloy.

8
Mater. Res. Express 6 (2019) 0865i1 L Guoliang et al

Figure 10. The Nyquist(a) and Bode curves(b) of Zn-1Mg alloy and β-TCP/Zn-1Mg composite in the SBF.

Figure 10 shows electrochemical impedance spectroscopy of Zn-1Mg alloy and β-TCP/Zn-1Mg composite
in SBF. Zn-1Mg alloy and β-TCP/Zn-1Mg composite exhibited similar the Nyquist and Bode curves indicating
similar electrochemical behavior. There was a capacitive arc in the high frequency region suggesting the charge
transfer resistance and the electric double layer capacitance on the alloy surface. The polarization resistance of
the Zn-1Mg alloy is 818Ω·cm2, and that of the β-TCP/Zn-1Mg composite is 634 Ω·cm2. The Zn-1Mg alloy has a
high polarization resistance, indicating that Zn-1Mg alloy has good electrochemical corrosion resistance. Bode
curves, as shown in figure 10(b), showed only one peak which corresponds to the capacitive arc in the Nyquist
diagram throughout the test frequency range. In addition, the impedance magnitude of the Zn-1Mg alloy and
the β-TCP/Zn-1Mg composite in high frequency region were almost coincide. Compare with Zn-1Mg alloy, the
β-TCP/Zn-1Mg composite has a lower impedance magnitude in low frequency region, thus the β-TCP/Zn-
1Mg composite has stronger activity and higher degradation rate under the open circuit potential, which was
consistent with the results of potentiodynamic polarization curve. Hence, the addition of β-TCP particles could
reduce the corrosion resistance of β-TCP/Zn-1Mg composites and accelerate the corrosion process.

2.4. Cytotoxicity test


Figure 11 showed the RGR% of mouse osteoblasts as a function of different concentrations of Zn-1Mg and β-
TCP/Zn-1Mg composite extracts for 1 day, 3 days and 5 days culture. The results showed that the RGR% for the
Zn-1Mg alloy and β-TCP/Zn-1Mg composite was above 80%, in spite of different culture time and different
extract concentration. These results indicated that the extracts of alloys and composites at different
concentrations have no obvious toxicity effect on the MC3T3-E1 cells. As the immersion time increased, the
RGR% for the Zn-1Mg alloy has decreased and when the concentration of the extract exceeds 50%, the RGR%
also decreases. The cytotoxicity grade were grade 1 for Zn-1Mg alloy and β-TCP/Zn-1Mg composite.

9
Mater. Res. Express 6 (2019) 0865i1 L Guoliang et al

Figure 11. Relative growth rate of MC3T3-E1 in different extract concentration by MTT (a) Zn-1Mg, (b) β-TCP/Zn-1Mg.

3. Conclusions

In this study, Zn-1Mg alloy and β-TCP/Zn-1Mg composites are fabricated using stirring casting assisted by
ultrasonic vibration. The microstructure, mechanical properties, corrosion behavior and cytotoxicity of the
alloy and composite were systematically investigated. The obtained results are summarized as follows:

(1) Compared with the Zn-1Mg alloy, β-TCP caused significantly grain refinement in the β-TCP/Zn-1Mg
composite.
(2) The UTS, Y.S. and elongation of β-TCP/Zn-1Mg composite is 330.1 MPa, 249.3 MPa and 11.7%,
respectively. Compared with Zn-1Mg alloy, the UTS, Y.S. and elongation of β-TCP/Zn-1Mg composite
increased by 9.9%, 10.2% and 101.7%, respectively.
(3) In-vitro immersion showed that the stable corrosion rate of β-TCP/Zn-1Mg composites in SBF solution
was 0.046 mm/a similar to that of the Zn-1Mg alloy.
(4) Compare to Zn-1Mg alloy, an increased corrosion rate was confirmed for β-TCP/Zn-1Mg composites by
electrochemical analysis.
(5) The RGR% for 1∼5 days culture of the osteoblasts with the extracts of Zn-1Mg alloy or β-TCP/Zn-1Mg
composites was above 80%, thus the cytotoxicity grade were grade 1.

Acknowledgments

The authors are grateful for the support from the National Nature Science Foundation of China (No.
U1764254), Innovative training program for college students (201810060033).

10
Mater. Res. Express 6 (2019) 0865i1 L Guoliang et al

ORCID iDs

Liu Debao https://orcid.org/0000-0003-0533-0693

References
[1] Zheng Y F, Gu X N and Witte F 2014 Biodegradable metals Materials Science and Engineering R Reports 77 1–34
[2] Chen Q and Thouas G A 2015 Metallic implant biomaterials Mater. Sci. Eng. R 87 1–57
[3] Li H, Zheng Y and Qin L 2014 Progress of biodegradable metals Progress in Natural Science: Materials International 24 414–22
[4] Witte F 2010 The history of biodegradable magnesium implants: a review Acta Biomater. 6 1680–92
[5] Staiger M P, Pietak A M, Huadmai J and Dias G 2006 Magnesium and its alloys as orthopedic biomaterials: a review Biomaterials 27
1728–34
[6] Liu D, Liu Y, Zhao Y et al 2017 The hot deformation behavior and microstructure evolution of HA/Mg-3Zn-0.8Zr composites for
biomedical application J. Materials Science and Engineering C 77 690–97
[7] Bowen P K, Drelich J and Goldman J 2013 Zinc exhibits ideal physiological corrosion behavior for bioabsorbable stents Adv. Mater. 25
2577–82
[8] Moonga B S and Dempster D W 2010 Zinc is a potent inhibitor of osteoclastic bone resorption in vitro J. Bone Min. Res. 10 453–7
[9] Vojtěch D et al 2011 Mechanical and corrosion properties of newly developed biodegradable Zn-based alloys for bone fixation Acta
Biomater. 7 3515–22
[10] Gong H, Wang K, Strich R and Zhou J G 2015 In vitro biodegradation behavior, mechanical properties, and cytotoxicity of
biodegradable Zn–Mg alloy J. Biomed. Mater. Res. B Appl. Biomater. 103 1632–40
[11] Kubásek J et al 2016 Microstructure and mechanical properties of the micrograined hypoeutectic Zn–Mg alloy Int. J. Miner. Metall.
Mater. 23 1167–76
[12] Liu D B et al 2012 Fabrication of biodegradable nano-sized β-TCP/Mg composite by a novel melt shearing technology Mater. Sci. Eng.
C 32 1253–8
[13] Serra I R et al 2015 Production and characterization of chitosan/gelatin/β-TCP scaffolds for improved bone tissue regeneration Mater.
Sci. Eng. C 55 592–604
[14] ASTM 2009 E8/E8M - 09 Standard Test Methods For Tension Testing of Metallic Materials 1–27
[15] ISO 10993-5: 2009 2009 Biological evaluation of medical devices, Part 5: tests for in vitro cytotoxicity Biol. Eval. Med. Devices. 1–34

11

You might also like