You are on page 1of 12

Cite this article Research Article Keywords: alloys/biodegradable/surface

Törne KB, Khan FA, Örnberg A and Weissenrieder J (2018) Paper 1700053 characterisation
Zn–Mg and Zn–Ag degradation mechanism under biologically relevant conditions. Received 20/09/2017; Accepted 01/11/2017
Surface Innovations 6(1–2): 81–92, Published online 21/11/2017
https://doi.org/10.1680/jsuin.17.00053
ICE Publishing: All rights reserved

Surface Innovations

Zn–Mg and Zn–Ag degradation mechanism


under biologically relevant conditions
Karin Beaussant Törne PhD Andreas Örnberg Lic, MSc
Material Physics, Department of Applied Physics, KTH Royal Institute of Abbott, St Paul, MN, USA
Technology, Stockholm, Sweden; Abbott, St Paul, MN, USA Jonas Weissenrieder PhD
Fareed Ashraf Khan PhD Associate Professor, Material Physics, Department of Applied Physics,
Department of Materials Science and Engineering, KTH Royal Institute of KTH Royal Institute of Technology, Stockholm, Sweden
Technology, Stockholm, Sweden (corresponding author: jonas@kth.se)

Zinc (Zn) alloys form a promising new class of biodegradable metals that combine suitable mechanical properties with
the favorable degradation properties of pure zinc. However, the current understanding of the influence of alloying
elements on the corrosion of zinc alloys, in biologically relevant media, is limited. The authors studied the degradation
of three alloys, zinc–4 wt% silver (Ag), zinc–0·5 wt% magnesium (Mg) and zinc–3 wt% magnesium by in situ
electrochemical impedance spectroscopy (EIS). After exposure for 1 h or 30 d, the samples were characterized by
infrared spectroscopy and scanning electron microscopy. The presence of secondary phases in the alloy microstructure
induced selective corrosion and increased the degradation rate. EIS analysis revealed an increase in surface
inhomogeneity already at short (hours) immersion times. The microgalvanic corrosion of the zinc–silver alloy resulted in
enrichment of the AgZn3 phase at the sample surface. The enrichment of silver and potential release of AgZn3 particles
may result in complications during the tissue regeneration. The zinc–magnesium alloy surface was depleted of the
magnesium-rich phase after 8–12 d. The selective dissolution caused local precipitation of corrosion products and a
thicker corrosion layer with larger pore size consistent with increased corrosion rate.

1. Introduction biocompatibility. Cytotoxicity studies have shown a reduced


For more than a decade, biodegradable metals have been proposed viability of vascular smooth muscle cells grown in zinc alloy
as a potential material for temporary implants.1 The combination of extracts.13,14 However, numerous in vivo studies show thin
a simple degradation profile and mechanical strength is a property thickness of neointimal tissue and tissue regeneration.4,13 Vojtech
that renders biodegradable metals attractive for future et al.11 estimated the daily release from a zinc bone screw to be
cardiovascular and orthopedic applications.2,3 Currently permanent 1·5 mg/d. This can be compared to the 2–3 g of zinc contained in
metallic implants are also used for applications where the implant is the human body and the recommended dietary allowance of
necessary only during a healing and remodeling phase. A 15 mg/d.11,15,16 The corrosion rate of pure zinc appears to be ideal
permanent implant may cause a late adverse effect and require a for cardiovascular stent and bone fixation applications in order to
second surgery to remove the implant.3,4 Degradable implants can retain the mechanical properties during the entire healing
reduce cost and more importantly improve the patient’s quality of process.11,12 Both in vivo and in vitro studies have determined the
life by avoiding explantation surgery.5,6 Iron and magnesium (Mg) corrosion rate to be in the order of tens of micrometers per year,
alloys were first suggested due to their low toxicity.1 Magnesium between the degradation rates of magnesium and iron.11,12,17 Pure
alloys also possess a favorable mechanical stiffness close to that of zinc, however, is soft and exhibit relatively low tensile strength. In
cortical bone.3 However, in vivo studies showed that the corrosion order to improve the mechanical properties of zinc, researches have
rate of magnesium is too fast for most applications.1,7 Further, suggested several alloying elements. Li et al.13 demonstrated that
magnesium degradation often produces copious amounts of magnesium, calcium (Ca) and strontium (Sr) are suitable candidates
hydrogen gas.1 Hydrogen gas bubbles formed in the vicinity of the that may result in alloys with increased strength and ductility
implant surface may complicate the healing process.7 Hydrogen compared to pure zinc. Further improvements were obtained with
may also be absorbed by the remaining magnesium implant, which ternary alloys zinc–magnesium–X, where X is calcium or
can result in hydrogen embrittlement and premature failure of load- strontium.18 More recently, a number of studies were published
bearing implants.8 Iron, on the other hand, degrades too slowly for suggesting zinc–magnesium, zinc–aluminum (Al), zinc–lithium (Li)
many applications and produces voluminous corrosion products, and zinc–silver (Ag) as potential biodegradable alloys.19–22 The
causing an increase in the volume of the implant.9,10 zinc–silver alloy was shown to possess high mechanical strength as
well as superplasticity.22 Current knowledge suggests zinc alloys to
Zinc (Zn) was recently suggested as an alternative degradable metal be a promising new class of biodegradable metals. However, there
or alloy base.11,12 The toxicity of zinc is higher than that of iron or is still a lack of understanding of what effect alloying has on the
magnesium, and concerns have therefore been raised regarding its corrosion mechanism and temporal evolution of the surface

81
Downloaded by [ Universidad Nacional Autonoma de México (UNAM)] on [17/09/18]. Copyright © ICE Publishing, all rights reserved.
Surface Innovations Zn–Mg and Zn–Ag degradation
Volume 6 Issue SI1–2 mechanism under biologically relevant
conditions
Törne, Khan, Örnberg and Weissenrieder

properties in a biologically relevant environment. It is of specific Table 1. Chemical compositions of cast alloy samples (as
interest to establish how the microstructure of the alloys influence determined by ICP-OES)
the interface properties of the corroding surface, at different times Alloy Magnesium: wt% Silver: wt% Zinc: wt%
in the degradation process, and how it may be at variance from that
Zn0·5Mg 0·62 ± 0·04 — Bal.
of pure zinc samples. In the current study, the authors employed Zn3Mg 2·6 ± 0·3 — Bal.
electrochemical techniques to study the initial (hours) and long- Zn4Ag — 4·4 ± 0·02 Bal.
term (30 d) degradation mechanism of two zinc–magnesium alloys
Bal., balance
and one zinc–silver alloy and compare the results to those for pure
zinc samples. Through the use of electrochemical impedance
spectroscopy (EIS), which allows for in situ characterization of the 2.2 Electrochemical characterization
corroding interface during immersion, new information on the Electrochemical analysis was performed with a VMP2
effect of alloying on the biodegradation of zinc may be obtained. potentiostat/galvanostat/frequency response analyzer (Bio-Logic
By characterizing the surface with EIS over a long term, it is Science Instruments). A three-electrode flat cell was used for all
possible to detect time-dependent changes in the interface and the measurements. The reference electrode was a silver/silver chloride
corrosion mechanism. Together with careful ex situ examination, (AgCl)-saturated potassium chloride (KCl) electrode. All
including scanning electron microscopy (SEM) of the surface and potentials are given relative to this reference electrode. The
of the cross-sections, a detailed description of the interfaces and counter electrode was a large-surface-area platinum (Pt) mesh.
corrosion mechanisms is provided.
Initial corrosion rates were obtained from potentiodynamic
polarization (PDP) scans, performed after 30 min immersion, from
2. Materials and methods
−20 to +250 mV against the open-circuit potential (OCP) at a scan
2.1 Material fabrication and characterization rate of 0·16 mV/s, according to ASTM G 5-14.26
The alloys were prepared by melting pure zinc and either
magnesium or silver in an alumina crucible in ambient A long-term (30 d) immersion study was performed in Ringer’s
atmosphere at 550°C for 20 min without stirring. The melt was solution at 37°C and pH 7·4. Three samples each of the alloys and
casted at room temperature into rods 12 mm dia. and 100 mm cast zinc were evaluated. Previous studies showed Ringer’s solution
long. The middle 80 mm were cut into samples for analysis; the to be a suitable model electrolyte for long-term in vitro studies of
top and bottom 10 mm of the rod were excluded from this study. zinc biodegradation.17 The corrosion of pure zinc in Ringer’s
Cast rods of pure zinc were prepared as reference samples by solution was homogeneous and similar to reported in vivo
using the same process. The targeted composition of the alloys results.27,28 Plasma and whole blood are preferred electrolytes for
was 0·5 and 3 wt% magnesium and 4 wt% silver, here denoted as short-term evaluations as they provide an environment closer to the
Zn0·5Mg, Zn3Mg and Zn4Ag, respectively. The compositions clinical situation.17 However, plasma and whole blood degrade with
were chosen as previous studies had shown them to be of time and are therefore unsuitable for long-term evaluations. Three
particular interest as degradable alloys. Alloys with 0·5 wt% of hundred milliliters of buffer was used per sample, and 132 ml was
magnesium have been shown to possess a suitable combination continuously exchanged, using a peristaltic pump, every 24 h to
of strength and elasticity, and 3 wt% is a eutectic point of the prevent changes in pH or increasing concentration of zinc ions
zinc–magnesium system.11,19,20 For the silver alloys, Liu et al.23 (Zn2+). The exposed surface area of the sample was 0·38 cm2. EIS
demonstrated that a minimal grain size is obtained by alloying impedance was measured at regular intervals during the immersion
with 4 wt% silver, which implies higher strength. time. A low perturbation amplitude of 2 mV and a frequency range
between 100 000 and 0·1 Hz was used to avoid inducing
Microstructural characterization was performed by optical irreversible changes to the sample surface, as determined in the
microscopy (Leica DMLM) and SEM (Carl Zeiss AG Ultra 55 authors’ previous study on zinc corrosion.17
with an Oxford Instruments Inca PentaFet x3 energy-dispersive
X-ray spectroscopy analyzer) on samples polished according to Following the 30 d immersion, the sample surface was
ASTM standard E 3-11.24 The final polishing was performed with characterized by SEM and the composition of the corrosion
colloidal silica, which also results in gentle etching of the samples product was determined by Fourier transform infrared
that reveals the microstructure. spectroscopy (FTIR) (Thermo Scientific Nicolet iS 10). After
initial examination of the sample surface, the corrosion product
The composition of the alloys was analyzed by inductively was removed from two of the three samples by cleaning in 12%
coupled plasma optical emission spectroscopy (ICP-OES) ammonia solution, according to ASTM G 1.29 The treatment
(Agilent 720). The average compositions and stand deviation are allowed investigation of the underlying metallic surface. Polished
presented in Table 1. samples were also immersed in the ammonia solution and
investigated by SEM to confirm that the cleaning process
The hardness of the samples was measured using a Vickers indenter removed only the corrosion product and did not affect the metallic
with 4·9 N (500 g-F) load, according to ASTM E 384-11.25 surface. The third sample was used to prepare cross-sections by

82
Downloaded by [ Universidad Nacional Autonoma de México (UNAM)] on [17/09/18]. Copyright © ICE Publishing, all rights reserved.
Surface Innovations Zn–Mg and Zn–Ag degradation
Volume 6 Issue SI1–2 mechanism under biologically relevant
conditions
Törne, Khan, Örnberg and Weissenrieder

mounting in epoxy resin and polishing down to the desired depth structure at the grain boundaries (dark areas) (Figure 1(c)). A
with silicon carbide (SiC) abrasive paper. Three cross-sections, high-resolution SEM image of the eutectic structure is included in
two at opposing edges and one in the middle, were examined per Figure S2 in the online supplementary material. The eutectic
sample. Additional samples were also immersed in Ringer’s structure has been assigned as consisting of zinc and Mg2Zn11.30
solution for 1 h for short-term microscopic evaluation of the Calculations using the level rule predicts the Zn0·5Mg alloy to be
surface structure. composed of 80 wt% primary zinc grains and 20 wt% eutectic
structure. The microstructure of the Zn3Mg alloy exhibited
mainly a lamellar eutectic structure (Figure 1(d)). Due to the
3. Results
deviation from the eutectic composition (2·6 wt% magnesium
3.1 Composition and microstructure rather than 3 wt%), primary zinc grains were observed.
The chemical compositions of the zinc alloys as determined by Calculations predict the Zn3Mg alloy to consist of 13 wt%
ICP-OES are presented in Table 1. The cast zinc samples exhibit primary grains. By similar calculations using the lever rule, the
grains of 20–50 mm wide (Figure 1(a)). Micropores formed during eutectic structure is composed of 57 wt% zinc and 43 wt%
the casting process were observed in the cast zinc samples as well Mg2Zn11. Microhardness analysis showed increasing hardness of
as the alloys (selected pores are indicated by arrows in Figure 1). the alloys relative to the cast zinc. The hardness of the
A high-resolution SEM image of a micropore is included in zinc–magnesium alloys increased with increasing magnesium
Figure S1 in the online supplementary material. The addition of content, and the Zn4Ag alloy had a hardness similar to that of the
the alloying elements resulted in significant grain refinement Zn0·5Mg alloy (Figure 2).
and formation of secondary phases. The Zn4Ag samples feature a
dendritic structure (Figure 1(b)) with dendrites 1–10 mm wide and 3.2 Potentiodynamic polarization
10–50 mm long. Liu et al.23 characterized the refinement of the PDP analysis of cast zinc and zinc alloy samples after 30 min
microstructure in zinc–silver alloys as due to the initial formation immersion in Ringer’s solution is shown in Figure 3(a). A
of the e-AgZn3 phase, which serves as a nucleation site for the h- decrease in the corrosion potential of the alloys compared to that
zinc phase in a peritectic solidification reaction. The insert in of the cast zinc was observed. Initial corrosion rates were
Figure 1(b) shows a high-magnification SEM image of a AgZn3 calculated from Tafel fits (Figure 3(b)). There was no statistically
particle.22,23 The Zn0·5Mg alloy microstructure was characterized significant difference between the cast zinc and the Zn0·5Mg
by primary zinc dendrites (light areas) with size of approximately alloy. A small, but statically significant, increase in corrosion rate
20 mm and precipitates of a magnesium-containing eutectic was observed for the Zn3Mg and Zn4Ag alloys. The Tafel fit

10 µm

100 µm 100 µm

(a) (b)

50 µm 50 µm

(c) (d)

Figure 1. Optical micrographs of (a) cast zinc sample with grain size of 20–50 mm and (b) Zn4Ag sample with dendritic grain structure.
Insert is a SEM image of the AgZn3 precipitates at high magnification. (c) Zn0·5Mg sample with h-zinc grains (bright areas) and eutectic
phase (dark areas) at grain boundaries. (d) Zn3Mg sample with high density of eutectic phase and only scattered zinc grains. The arrows
indicate selected microporous defects from the casting procedure

83
Downloaded by [ Universidad Nacional Autonoma de México (UNAM)] on [17/09/18]. Copyright © ICE Publishing, all rights reserved.
Surface Innovations Zn–Mg and Zn–Ag degradation
Volume 6 Issue SI1–2 mechanism under biologically relevant
conditions
Törne, Khan, Örnberg and Weissenrieder

immersion times, no difference in OCP could be observed


140
between zinc and the zinc alloys.
120
3.4 Electrochemical impedance spectroscopy
Microhardness: HV

100
Representative Bode plots of the impedance response of the
80 interface at selected time points during the 30 d immersion of the
cast zinc and zinc alloy samples are presented in Figures 5(a)–5(d).
60
The impedance at short immersion time of all samples shows two
40 apparent time constants. For the alloy samples, the first time
constant appears at ~100 Hz and the second at 0·1–1 Hz.
20
However, the low-frequency time constant of zinc was observed
0 at slightly higher frequencies, around 10 Hz (Figure 5(a)). The
Zn Zn0·5Mg Zn3Mg Zn4Ag
high- and low-frequency time constants have previously been
Figure 2. Vickers hardness of cast zinc (35 ± 5·5 HV) and zinc ascribed to the charge transfer process and a diffusion process
alloys, Zn0·5Mg (65 ± 5·4 HV), Zn3Mg (127 ± 8·2 HV) and Zn4Ag through a porous layer of corrosion product.17,31 For most of the
(66 ± 3·2 HV). Error bars indicates one standard deviation of seven samples, the low-frequency time constant was absent at long
measurements over three samples

(a)
−0·96
analysis from each PDP scan is included in Tables S1–S4 in the −1·04
OCP against Ag/AgCI: V

online supplementary material.


(b)
−0·96
3.3 Open-circuit potential −1·04
The development of the OCP during the 30 d immersion study is
−0·96 (c)
presented in Figures 4(a)–4(d). The initial OCP of the Zn0·5Mg
and Zn3Mg alloys was shifted cathodically compared to that of −1·04
the cast zinc. This shift is expected as magnesium has a more (d)
−0·96
cathodic OCP than zinc. The Mg2Zn11 phase is therefore expected
to have a more cathodic potential than the zinc phase. The initial −1·04
OCP of the Zn4Ag alloy was also shifted cathodically compared 0 5 10 15 20 25 30
to that of zinc during the first hour of immersion but increased Time: d
rapidly to values similar to that of zinc (not shown). During the
Figure 4. Development of OCP during immersion of samples in
immersion period, the OCP of all samples increased anodically Ringer’s solution over 30 d: (a) cast zinc; (b) Zn4Ag; (c) Zn3Mg;
from between −1·06 and −1·03 V to between −0·99 and −0·93 V, (d) Zn0·5Mg
suggesting the buildup of a protective corrosion layer. At long

−0·85
0·20
−0·90
Potential against Ag/AgCl: V

Corrosion rate: mm/year

0·15
−0·95

−1·00 0·10

−1·05 0·05

−1·10
0
1 × 10−7 1 × 10−6 1 × 10−5 1 × 10−4 1 × 10−3 0·01 Zn Zn0·5Mg Zn3Mg ZnAg
Current density: A/cm (b)
(a)

Figure 3. (a) PDP analysis of zinc and zinc alloys in Ringer’s solution after 30 min immersion. The black solid line is pure cast zinc; the red
dashed line, Zn4Ag; the green dotted line, Zn3Mg; and the blue dashed dotted line, Zn0·5Mg. (b) Corrosion rates obtained by Tafel fit of
the PDP analysis. Error bars indicate one standard deviation of three samples

84
Downloaded by [ Universidad Nacional Autonoma de México (UNAM)] on [17/09/18]. Copyright © ICE Publishing, all rights reserved.
Surface Innovations Zn–Mg and Zn–Ag degradation
Volume 6 Issue SI1–2 mechanism under biologically relevant
conditions
Törne, Khan, Örnberg and Weissenrieder

90 90
103 103
Absolute impedance: W/cm2

Absolute impedance: W/cm2


70 70

Phase angle

Phase angle
50 50
102 102
30 30

10 10
10 10
0·1 1 10 100 1000 10 000 100 000 0·1 1 10 100 1000 10 000 100 000
Frequency: Hz Frequency: Hz
(a) (b)
90 90
103 103
Absolute impedance: W/cm2

Absolute impedance: W/cm2


70 70
Phase angle

Phase angle
50 50
102 102
30 30

10 10
10 10
0·1 1 10 100 1000 10 000 100 000 0·1 1 10 100 1000 10 000 100 000
Frequency: Hz Frequency: Hz
(c) (d)

Figure 5. EIS Bode plots of zinc alloys during 30 d immersion in Ringer’s solution: (a) pure cast zinc; (b) Zn4Ag; (c) Zn0·5Mg; (d) Zn3Mg.
■, 1 h; ●,10 d; ▲, 20 d; ▼, 30 d. Filled symbols are the absolute impedance, empty symbols the phase angle and solid lines indicate
fitted data

immersion times (>12 d), indicating the breakdown of the growing corrosion layer. After a few days of immersion, the
protective corrosion product layer and the onset of localized trend reversed and the capacitance increased, from 20 to
corrosion. The time point at which the change in corrosion 100 msh/(W cm2). The increase may potentially be due to increasing
mechanism occurred seems stochastic, with no apparent trend surface roughness. The capacitance of the film of corrosion
between the alloys. For some samples (one cast zinc, one products on cast zinc samples was roughly one order of magnitude
Zn0·5Mg and one Zn3Mg), the low-frequency time constant (~10–100 msh/(W cm2)) lower compared to that for the alloy
remained present throughout the entire immersion period. This samples (~400 msh/(W cm2)) at all time points. The exponential
suggests that the breakdown of the corrosion product layer did not term h of the CPE was 0·7–0·9 for both cast zinc and the alloys.
yet occur for these samples.

The impedance data with two time constants were fitted to the
CPEF
equivalent circuit in Figure 6(a) using the ZSimpWin 3.21 software. CPEDL
Constant-phase elements were used rather than ideal capacitors to
RS
compensate for surface inhomogeneity.32 CPEF describes the RS CPEP
capacitance of the corrosion film. Rp represents the resistance of the RCT
Rp
pores. CPEp describes the capacitance of the metal solution interface
RCT (b)
at the pore bottom. Rct describes the charge transfer resistance. The
spectra consisting of one time constant were fitted using a Randles- (a)
type equivalent circuit (Figure 6(b)) CPEDL describes the double
layer capacitance. All fitted data are included in Tables S5–S16 in the Figure 6. Equivalent circuits used to simulate the EIS data in
online supplementary material. ZSimpWin 3.21. (a) Circuit describing a porous interface. Used for
samples at short and medium immersion times < ~12 d.
(b) Randles circuit describing the surface after breakdown of the
The capacitance of the corrosion film on cast zinc samples initially porous layer. Used at long immersion times > ~12 d
decreased with time, from 60 to 20 msh/(W cm2), suggesting a

85
Downloaded by [ Universidad Nacional Autonoma de México (UNAM)] on [17/09/18]. Copyright © ICE Publishing, all rights reserved.
Surface Innovations Zn–Mg and Zn–Ag degradation
Volume 6 Issue SI1–2 mechanism under biologically relevant
conditions
Törne, Khan, Örnberg and Weissenrieder

The pore resistance of cast zinc increased over the first days from exponential term h was determined to be ~0·5, consistent with a
~30–100 to ~500–1000 W/cm2 and stabilized thereafter. However, diffusion-controlled process.
large fluctuations with time were observed. Increasing pore resistance
suggests growing pore length – that is, increasing film thickness. The The resistance related to the charge transfer resistance Rct of cast zinc
fluctuations with time can be explained by cycles of formation and increased during the first days to ~1000 W/cm2, suggesting increasing
detachment of the corrosion layer. The Rp of the alloys was lower protection with time. The Rct of Zn0·5Mg decreased initially to
compared to that of cast zinc, suggesting a thinner layer or increased 20–70 W/cm2 after 2–8 d and then increased to reach ~400 W/cm2
pore size. The Zn4Ag alloy stabilized at 100 W/cm2 and occasionally after 10 d. The Rp of Zn3Mg decreased to 150 W/cm2 after 2–8 d and
increased to ~400 W/cm2. In a similar way, the Rp of the Zn0·5Mg subsequently increased to around 500 W/cm2. The Rp of Zn4Ag
alloy stabilized at 40–200 W/cm2. Intermittently, the Rp increased to remained around 200 W/cm2 for the majority of the immersion time.
400 W/cm2, a behavior consistent with cycles of formation and
detachment of the corrosion layer. The Zn3Mg alloy demonstrated 3.5 Ex situ examination
the lowest Rp, which remained below 100 W/cm2. 3.5.1 Fourier transform infrared spectroscopy
The vibration modes present in the FTIR spectra obtained from
Similar to film capacitance, the CPEp for cast zinc samples the alloys after immersion did not differ from cast zinc, indicating
stabilized at lower values (~300–1000 msh/(W cm2)) compared to a similar chemical composition of the majority of the corrosion
the zinc alloys (10 000–20 000 msh/(W cm2)). The increased product of all samples, a hydrated carbonate. Figure 7 shows a
capacitance of the alloys is consistent with larger pore size and representative spectrum obtained from a Zn0·5Mg sample. The
with higher porosity, corresponding to a larger surface area. The strong peak at 3350 cm−1 corresponds to H2O stretching
vibrations. The four bands at 1550, 1490, 1420 and 1390 cm−1 are
assigned to antisymmetric v3 CO32− modes. The bands at 1040
and 940 cm−1 correspond to OH liberation modes. The sharp peak
at 820 cm−1 is ascribed to the out-of-plane CO32− bending mode.
The two bands at 730 and 690 cm−1 correspond to the v4 CO32−
Transmittance: %

modes. The spectra are overall in good agreement with literature


spectra of hydrozincite.33
3350

940
3.5.2 SEM after 1 h immersion
1550
1390 1040 SEM images of zinc samples exposed for 1 h showed a thin layer
1420 of corrosion product precipitated onto the surface (Figure 8(a)).
1490 730
The surface of the alloys showed a non-uniform layer of corrosion
820690 product of similar structure but with pores (pore size: 1–10 mm)
4000 3500 3000 1500 1000 scattered over the surface. Figure 8(b) shows an image of
Wave number: cm−1 representative pores on a Zn3Mg sample, indicated by the
encircled areas. The dashed line surrounds the agglomerated
Figure 7. FTIR spectra of corrosion product on a Zn0·5Mg sample. corrosion product around one of the pits. The porous structure of
The spectra obtained from cast zinc and the other alloys exhibit
the same modes, and the overall spectral shape agrees well with
the corrosion layer on Zn0·5Mg and Zn4Ag were similar to the
hydrozincite (Zn5(CO3)2(OH)6)33 corrosion layer on Zn3Mg with SEM images included in
Figure S3 in the online supplementary material.

4 µm 4 µm

(a) (b)

Figure 8. SEM images of the corrosion product layer after 1 h exposure: (a) cast zinc with a homogeneous layer of corrosion products;
(b) Zn3Mg alloys exhibiting micropores in the corrosion layer, indicated as encircled areas, and agglomerated corrosion product, indicated
by the dashed line. Similar pores and product were detected on the other alloy surfaces as well

86
Downloaded by [ Universidad Nacional Autonoma de México (UNAM)] on [17/09/18]. Copyright © ICE Publishing, all rights reserved.
Surface Innovations Zn–Mg and Zn–Ag degradation
Volume 6 Issue SI1–2 mechanism under biologically relevant
conditions
Törne, Khan, Örnberg and Weissenrieder

3.5.3 SEM of cast zinc (30 d) visible in the surface structure, as indicated by the dashed lines in
SEM images of the surface of cast zinc samples after 30 d of Figure 9(b). EDS analysis detected low levels of carbon and
immersion showed a thick compact layer of corrosion product oxygen and high levels of zinc and silver, likely from the
(Figure 9(a)). A structure with a similar appearance as the grain underlying metal. After the corrosion product was removed, the
microstructure with a thinner corrosion product at the grain surface structure demonstrated clear signs of microgalvanic
boundaries and thicker toward the center of the grains was corrosion with selective corrosion surrounding the AgZn3 particles
observed. This structure suggests increased anodic oxidation at (Figure 10(b)). The selective anodic dissolution surrounding the
the grain boundaries and cathodic reduction and subsequent AgZn3 dendrites are indicated by dashed lines encircling the
corrosion product precipitation at the grain center. In agreement attack. However, there was no sign of the large corrosion pits
with the FTIR results, EDS analysis confirmed the presence of (sized in millimeters) observed on the cast zinc samples. The
carbon (C), oxygen (O) and zinc. EDS results from all alloys are cross-sectional analysis showed a highly inhomogeneous surface
included in Figure S5 in the online supplementary material. Some with pores penetrating tens of micrometers down the metal matrix
areas of the corrosion film exhibited millimeter-sized defects and (Figure 11(b)). There was no detectable corrosion layer on the
exposed deep pits, confirming the film breakdown suggested by areas which were not covered by the local thick precipitation of
the EIS results. Similar large defects were observed on the alloy corrosion product. This suggests that any precipitated corrosion
samples as well, in contrast to the microscopic defects observed layer on these areas is thinner than 1 mm.
after 1 h, which were observed only on the alloys. After the
corrosion product was removed, SEM images showed that 3.5.5 SEM of Zn0·5Mg and Zn3Mg (30 d)
the majority of the surface had a relatively smooth structure The surface of Zn0·5Mg samples after 30 d immersion showed a
(Figure 10(a)). Deep (hundreds of micrometers) pits, 1–2 mm porous corrosion product; the pore size was <1 mm with local
wide, were also observed, confirming the presence of localized precipitates forming a denser corrosion product (Figure 9(c)). The
corrosion (insert in Figure 10). A cross-sectional SEM image of a insert is a magnification of the porous structure. EDS detected
pit is included in Figure S4 in the online supplementary material. carbon, oxygen and zinc with no difference in elemental
Cross-sectional SEM images revealed a smooth homogenous composition between the dense and porous corrosion products.
coating ~10 mm thick (Figure 11(a)). The surface of Zn3Mg was covered by clusters of corrosion
product, which lump together to form ridges and pores with a size
3.5.4 SEM of Zn4Ag (30 d) of few micrometers (Figure 9(d)). Millimeter-wide pits due to
Parts of the surface of the Zn4Ag samples were covered by a localized corrosion similar to what was reported earlier for cast
thick corrosion product, shown in the inset in Figure 9(b). The zinc samples were observed on both alloys. The cleaned samples’
majority of the surface was, however, covered only by a thin layer surface had a layered structure similar to that of the cast zinc
of corrosion product (Figure 9(b)). AgZn3 particles were clearly samples (Figures 10(c) and 10(d)). Small pits, <1 mm, indicated

80 µm

20 µm 20 µm

(a) (b)

1 µm

20 µm 20 µm

(c) (d)

Figure 9. SEM images of sample surface after 30 d of immersion. (a) Cast zinc with dense corrosion product. (b) Zn4Ag exhibiting a thin
corrosion film; the dashed line indicates the AgZn3 phase. Insert is at lower magnification showing the surface partly covered by thick
globular corrosion product. (c) Zn0·5Mg; insert is at higher magnification showing the microporous structure. (d) Zn3Mg with stratified
corrosion product

87
Downloaded by [ Universidad Nacional Autonoma de México (UNAM)] on [17/09/18]. Copyright © ICE Publishing, all rights reserved.
Surface Innovations Zn–Mg and Zn–Ag degradation
Volume 6 Issue SI1–2 mechanism under biologically relevant
conditions
Törne, Khan, Örnberg and Weissenrieder

1 mm

20 µm 20 µm

(a) (b)

20 µm 20 µm

(c) (d)

Figure 10. SEM images of sample surfaces after removal of corrosion product. (a) Cast zinc; insert is at lower magnification showing
localized corrosion attacks. (b) Zn4Ag; the dashed circles surrounding the preserved AgZn3 dendrites indicate areas significantly affected
by microgalvanic corrosion. (c) Zn0·5Mg and (d) Zn3Mg with a layered morphology depleted of eutectic phase; the arrows indicate
selected micropores

20 µm 20 µm

(a) (b)

20 µm 20 µm

(c) (d)

Figure 11. SEM images of interface cross-sections. (a) Cast zinc with a dense homogeneous layer of corrosion products approximately
10 mm thick. The white dashed line indicates the interface between the corrosion layer and the epoxy resin. (b) Zn4Ag; the dashed line
circles two AgZn3 particles with the surrounding matrix completely removed by the penetration of the anodic corrosion. (c) Zn0·5Mg with
pits filled with corrosion product 10 mm deep into the substrate. The dashed line circles one pit. (d) Zn3Mg with highly porous corrosion
layer, ~20 mm thick. The dashed line indicates the interface between the metal and the corrosion product

88
Downloaded by [ Universidad Nacional Autonoma de México (UNAM)] on [17/09/18]. Copyright © ICE Publishing, all rights reserved.
Surface Innovations Zn–Mg and Zn–Ag degradation
Volume 6 Issue SI1–2 mechanism under biologically relevant
conditions
Törne, Khan, Örnberg and Weissenrieder

by arrows, were observed, which were not observed on the cast O2 HO−
zinc samples. The EDS analysis on the cleaned surfaces of both Zn2+ + CO32−
alloys detected only very low levels of magnesium on the sample
surface. This magnesium depletion suggests a selective Zn2+
Zn5(CO3)2(OH)6
dissolution of the magnesium-rich phases. The cross-sections of AgZn3
Zn0·5Mg demonstrated an uneven surface with microcrevices Zn
filled with corrosion product ~30 mm deep (Figure 11(c)). The (a)
cross-sections of Zn3Mg also partly demonstrated a thick Mg2+ Zn2+
corrosion layer as well as some areas covered by a thinner more O2 HO−
irregular corrosion product as seen in Figure 11(d). Zn5(CO3)2
Zn Zn–Mg
4. Discussion
In agreement with previously reported results, the introduction of
alloying elements (magnesium and silver) to zinc increases the
hardness of the metal by both microstructure refinement and the
introduction of secondary phases.20 Increased hardness suggests Zn5(CO3)2
increased strength, and alloying appears to be a possible route to Zn
reach the desired mechanical properties of a biodegradable (b)
metallic implant.4 However, the presence of secondary phases
causes selective or microgalvanic corrosion due to potential Figure 12. Illustration of the corrosion mechanism of (a) Zn4Ag
differences between the primary zinc phase and the silver-/ with selective dissolution of the zinc phase surrounding the AgZn3
phase. Increased precipitation of corrosion products will occur on
magnesium-rich phases. The selective corrosion mechanism may areas with higher surface density of AgZn3 with an overall
result in deterioration of mechanical properties and more anodic potential and increased cathodic reaction
biocompatibility of an implant during its degradation. rates. (b) Zinc–magnesium alloy with selective dissolution of
the magnesium-rich phase. Increased precipitation and
The selective corrosion of the zinc phase in the Zn4Ag alloys is agglomeration of corrosion products on cathodic areas at short
immersion times (hours) (b, top image). At longer immersion
directly observed in the SEM analysis, where the AgZn3 dendrites times (days), most of the magnesium-rich phase is dissolved.
function as cathodic sites and the surrounding areas suffer Inclusions of denser corrosion product are formed on the
increased anodic dissolution (Figures 9–11). The local earlier cathodic sites of increased precipitation (b, lower image)
precipitation of corrosion product seen in the insert in Figure 9(b)
can be linked to an uneven distribution of AgZn3 particles from
the casting process. The corrosion mechanism is schematically
illustrated in Figure 12, where more corrosion product is expected SCC would hence also be of interest when further evaluating this
to precipitate on cathodic areas with denser population of AgZn3 new class of biodegradable alloys.
particles. The selective dissolution of zinc will result in an
enrichment of the silver-rich phase at the implant surface. As the The SEM images of the zinc–magnesium alloys also suggested
corrosion proceeds, AgZn3 particles may potentially be released microgalvanic corrosion (Figures 8–11). A schematic illustration
from the surface. Such change in interface properties should be of the corrosion mechanism is provided in Figure 12(b). Unlike
considered when evaluating the cyto- and biocompatibility of Zn4Ag, the magnesium-rich secondary phase is subjected to
zinc–silver alloys. Silver in itself is cytotoxic, but low amounts of anodic dissolution and the surrounding zinc phase constitutes
silver as an alloying element have been suggested to be beneficial the cathode areas. This results in selective dissolution of the
for imparting antibacterial properties to implants.34 Enrichment of magnesium-rich phase and removal of magnesium from the
silver on the surface with time may, however, shift the initially sample surface. The rate of dissolution and release of zinc and
beneficial antibacterial properties toward cytotoxic properties as magnesium (Mg2+) ions increase locally at the anodic areas. This
the silver concentration increases. Such changes may complicate can be observed in Figure 8(b), where pores form at anodic areas
the healing process. The initial fraction of AgZn3 is estimated to and precipitation of corrosion product occurs on the surrounding
be 3 vol% (from an initial composition of 4 wt% silver). As the cathodic areas. The micrometer-sized pits observed on the cleaned
implant degrades, the fraction of AgZn3 will increase. Future samples (Figures 10(c) and 10(d)) are of similar size and may be
studies on the biocompatibility and cytotoxicity on pure AgZn3 is traces after magnesium-rich particles. Surrounding the anodic
therefore of interest. The increasing surface area should also be areas are cathodic areas with increased precipitation of corrosion
considered as it may increase the amount of zinc released per unit product. The local precipitation and agglomeration of
time, relative to a pure zinc implant, and may reduce the life span corrosion product as observed in Figure 8(b) will form a denser
of the implant. The selective dissolution of magnesium in corrosion product at longer immersion times as observed on
magnesium alloys is associated with stress corrosion cracking Zn0·5Mg in Figure 9(c). The microstructure of Zn3Mg is mainly
(SCC).8 An evaluation of the sensitivity of zinc–silver alloys to eutectic (Figure 1(d)), with a very even and fine dispersion of zinc

89
Downloaded by [ Universidad Nacional Autonoma de México (UNAM)] on [17/09/18]. Copyright © ICE Publishing, all rights reserved.
Surface Innovations Zn–Mg and Zn–Ag degradation
Volume 6 Issue SI1–2 mechanism under biologically relevant
conditions
Törne, Khan, Örnberg and Weissenrieder

and Mg2Zn11. The fine distribution of cathodic sites (zinc) and demonstrated micropores that may be linked to the increased
local precipitation of corrosion product cause the stratified capacitance. At long immersion times, the high capacitance of
structure of the corrosion product (Figure 9(d)). Zn4Ag is, however, better rationalized by the large increase in
surface area due to the selective dissolution. The impedance results
The initial corrosion rate of Zn4Ag was increased compared to demonstrate that after 1 h of exposure, the selective corrosion of the
cast zinc, in agreement with the observed microgalvanic corrosion alloys already produces a less homogenous and more reactive
(Figure 3). These results support the previously reported increase surface compared to the cast zinc samples.
in corrosion rate with increasing silver content by Sikora-Jasinska
et al.22 The alloying of zinc with low amounts (0·5 wt%) of The initial charge transfer resistance Rct and pore resistance Rp of
magnesium did shift the initial OCP anodically, suggesting all the alloys were similar to those of cast zinc. However, Rct and
increased surface reactivity, but no significant increased corrosion Rp decreased over the first days to reach a minimum after 2–8 d
rate was observed in the PDP measurements. However, when the and stabilized at lower values compared to those of cast zinc at
amount of magnesium was increased to 3 wt%, an increase in long immersion times. The initial decrease is assigned to the in-
initial corrosion rate similar to that of Zn4Ag was observed. microscopy observed increase in surface area as well as the
Previous studies have reported decreased corrosion rates inhomogeneity of the surface from the selective dissolution. The
for Zn3Mg alloys due to the increased stability of the oxide increase in the Rct of zinc–magnesium after 12 d may be
layer.20,35 However, these studies were performed in phosphate rationalized as the complete depletion of the magnesium-rich
(PO43−)-containing solutions, which may result in decreased phase at the surface – that is, only very limited amounts remain
corrosion rates due to rapid formation of a passivating thin film after ~12 d. The lower resistance compared to that of cast zinc can
(nanometers thick) of phosphates.17,36 The passivating thin film also be explained by an increase in real surface area. However, Rct
causes localized corrosion at sites of breakdown while leaving the is a measure of the resistance to oxidation from the metallic state
majority of the surface virtually unaffected by the corrosion. The to the ionic state such that a decrease in Rct will be consistent
observed corrosion of pure zinc in in vivo studies is relatively with increased oxidation due to the microgalvanic corrosion
uniform with some localized corrosion, similar to the corrosion of induced by the potential difference between the two phases
pure zinc observed in this study.28 Although precipitation of present at the surface. The microgalvanic corrosion and
phosphate corrosion products is known to occur, in vivo these decreasing Rp are consistent with the observed thin corrosion
films are thicker (micrometers to millimeters) and less protective layer on the Zn4Ag alloy (Figure 9(b)) and increased porosity in
than the films formed in phosphate-containing solutions.17,37,38 the zinc–magnesium alloys (Figures 9(c) and 9(d)).
Therefore, corrosion rate assessment in phosphate solutions may
not be considered as representative. Ringer’s solution, which was 5. Conclusions
used in this study, does not contain PO42− and therefore provides The corrosion properties of zinc alloys, Zn4Ag, Zn0·5Mg and
a better model of the in vivo situation.17 Zn3Mg, were evaluated by electrochemical techniques, PDP and
EIS and ex situ SEM and FTIR. An increase in the corrosion rate
The impedance measurements of both cast zinc and the alloys of the alloys compared to cast zinc was observed in the PDP
indicated the buildup and breakdown of a protective layer of measurements. The impedance analysis suggested the presence of
corrosion products (Figure 5). The breakdown is indicated by the a porous corrosion film influencing the corrosion of the alloys.
absence of a low-frequency time constant at long immersion times, After 1 h of immersion, the impedance results show the surface
which occurred after 12–30 d. The interpretation agrees with the of alloy samples to be less homogeneous than that of the cast zinc
localized corrosion observed in the SEM images, as seen in the samples. At long immersion times, the Rct of the alloys were
insert in Figure 10(a). The breakdown may be initiated by local lower compared to that of cast zinc, suggesting continued
inhomogeneities such as the microspores observed in the increased corrosion. The lower Rct is consistent with the
microstructure (Figure 1). Further processing of the alloy by, for microgalvanic corrosion observed in the SEM analysis. The
example, extrusion can homogenize the structure and minimize microgalvanic corrosion of Zn4Ag caused a selective dissolution
defects, which would postpone the breakdown of the corrosion layer. of zinc and enrichment of AgZn3 at the sample surface during
immersion. Such enrichment of silver at the interface may induce
The analysis of the impedance data revealed increased capacitance problematic cytotoxic effects in a clinical situation and influence
for the alloys compared to cast zinc samples. Similar high the tissue regeneration process. Future studies of the
capacitances have previously been reported for zinc–aluminum–rare biocompatibility of AgZn3 would therefore be of interest for
earth alloys.31 Such high capacitances are typically associated with assessing the suitability of zinc–silver-based degradable metals.
microporous surface structures and large surface areas. Micropores Selective dissolution of magnesium-based alloys is known to
were observed on the alloy surfaces already after 1 h of exposure induce SCC. Investigations of the sensitivity of zinc–silver alloys to
(Figure 8(b)). At long immersion times, the microporous structure SCC would therefore also be of importance. The zinc–magnesium
of the corrosion product on the zinc–magnesium alloys (Figures alloys, on the other hand, demonstrated a depletion of magnesium
9(c) and 9(d)) is clearly visible and validates the impedance at the surface after ~12 d and increased corrosion resistance.
analysis. The surface of zinc–silver after 1 h exposure also Although the effect of alloying on the average corrosion rate was

90
Downloaded by [ Universidad Nacional Autonoma de México (UNAM)] on [17/09/18]. Copyright © ICE Publishing, all rights reserved.
Surface Innovations Zn–Mg and Zn–Ag degradation
Volume 6 Issue SI1–2 mechanism under biologically relevant
conditions
Törne, Khan, Örnberg and Weissenrieder

small, the microgalvanic corrosion induces significant local 16. Fosmire GJ (1990) Zinc toxicity. American Journal of Clinical
corrosion that may result in premature failure of the implant. The Nutrition 51(2): 225–227.
17. Törne K, Larsson M, Norlin A and Weissenrieder J (2016) Degradation
change in surface composition over time (due to the selective
of zinc in saline solutions, plasma, and whole blood. Journal of
dissolution) raises concerns for the biocompatibility. Such effects Biomedical Materials Research Part B: Applied Biomaterials 104(6):
should be considered when designing new materials, and the 1141–1151.
presence of secondary phases should be minimized. 18. Liu XW, Sun JK, Qiu KJ et al. (2016) Effects of alloying elements (Ca
and Sr) on microstructure, mechanical property and in vitro corrosion
behavior of biodegradable Zn–1.5Mg alloy. Journal of Alloys and
Acknowledgements
Compounds 664: 444–452.
Abbot and the Swedish Research Council (Vetenskapsrådet) are 19. Kubasek J, Vojtech D, Jablonska E et al. (2016) Structure, mechanical
acknowledged for their financial support. The authors would also characteristics and in vitro degradation, cytotoxicity, genotoxicity and
like to acknowledge David Chmielewski for assistance with the mutagenicity of novel biodegradable Zn–Mg alloys. Materials Science
microhardness test and Nathan Freshour for assistance with the ICP- & Engineering: C – Materials for Biological Applications 58: 24–35.
20. Mostaed E, Sikora-Jasinska M, Mostaed A et al. (2016) Novel Zn-
OES measurements. Hasse Fredriksson is gratefully acknowledged
based alloys for biodegradable stent applications: design, development
for assistance with the casting process. and in vitro degradation. Journal of the Mechanical Behavior of
Biomedical Materials 60: 581–602.
REFERENCES 21. Zhao S, Seitz JM, Eifler R et al. (2017) Zn–Li alloy after extrusion and
1. Zheng YF, Gu XN and Witte F (2014) Biodegradable metals. drawing: structural, mechanical characterization, and biodegradation
Materials Science and Engineering R: Reports 77: 1–34. in abdominal aorta of rat. Materials Science & Engineering: C –
2. Bowen PK, Drelich J and Goldman J (2013) A new in vitro–in vivo Materials for Biological Applications 76: 301–312.
correlation for bioabsorbable magnesium stents from mechanical 22. Sikora-Jasinska M, Mostaed E, Mostaed A et al. (2017) Fabrication,
behavior. Materials Science & Engineering: C – Materials for mechanical properties and in vitro degradation behavior of newly
Biological Applications 33(8): 5064–5070. developed ZnAg alloys for degradable implant applications. Materials
3. Staiger MP, Pietak AM, Huadmai J and Dias G (2006) Magnesium Science & Engineering: C – Materials for Biological Applications
and its alloys as orthopedic biomaterials: a review. Biomaterials 77: 1170–1181.
27(9): 1728–1734. 23. Liu ZL, Qiu D, Wang F, Taylor JA and Zhang MX (2014) The grain
4. Bowen PK, Shearier ER, Zhao S et al. (2016) Biodegradable metals for refining mechanism of cast zinc through silver inoculation. Acta
cardiovascular stents: from clinical concerns to recent Zn-alloys. Materialia 79: 315–326.
Advanced Healthcare Materials 5(10): 1121–1140. 24. ASTM (2017) E3-11: Standard guide for preparation of metallographic
5. Hofmann GO (1995) Biodegradable implants in traumatology – a specimens. ASTM International, West Conshohocken, PA, USA.
review on the state-of-the-art. Archives of Orthopaedic and Trauma 25. ASTM (2017) E3-84: Standard test method for microindentation
Surgery 114(3): 123–132. hardness of materials. ASTM International, West Conshohocken,
6. Peuster M, Beerbaum P, Bach FW and Hauser H (2006) Are PA, USA.
resorbable implants about to become a reality? Cardiology in the 26. ASTM (2014) ASTM G 5-14: Standard reference test method for
Young 16(2): 107–116. making potentiodynamic anodic polarization measurements. ASTM
7. Witte F, Hort N, Vogt C et al. (2008) Degradable biomaterials based International, West Conshohocken, PA, USA.
on magnesium corrosion. Current Opinion in Solid State & Materials 27. Torne K, Larsson M, Norlin A and Weissenrieder J (2016) Degradation
Science 12(5–6): 63–72. of zinc in saline solutions, plasma, and whole blood. Journal of
8. Winzer N, Atrens A, Song GL et al. (2005) A critical review of the Biomedical Materials Research Part B – Applied Biomaterials 104(6):
stress corrosion cracking (SCC) of magnesium alloys. Advanced 1141–1151.
Engineering Materials 7(8): 659–693. 28. Drelich AJ, Bowen PK, LaLonde L, Goldman J and Drelich JW (2016)
9. Pierson D, Edick J, Tauscher A et al. (2012) A simplified in vivo Importance of oxide film in endovascular biodegradable zinc stents.
approach for evaluating the bioabsorbable behavior of candidate stent Surface Innovations 4(3): 133–140.
materials. Journal of Biomedical Materials Research Part B – 29. ASTM (2011) G 1: standard practice for preparing, cleaning, and
Applied Biomaterials 100B(1): 58–67. evaluating corrosion test specimens. ASTM International, West
10. Peuster M, Hesse C, Schloo T et al. (2006) Long-term biocompatibility Conshohocken, PA, USA.
of a corrodible peripheral iron stent in the porcine descending aorta. 30. Prosek T, Nazarov A, Bexell U, Thierry D and Serak J (2008)
Biomaterials 27(28): 4955–4962. Corrosion mechanism of model zinc-magnesium alloys in atmospheric
11. Vojtech D, Kubasek J, Serak J and Novak P (2011) Mechanical conditions. Corrosion Science 50(8): 2216–2231.
and corrosion properties of newly developed biodegradable 31. Rosalbino F, Angelini E, Maccio D, Saccone A and Delfino S (2009)
Zn-based alloys for bone fixation. Acta Biomaterialia 7(9): Application of EIS to assess the effect of rare earths small addition on
3515–3522. the corrosion behaviour of Zn–5% Al (Galfan) alloy in neutral aerated
12. Bowen PK, Drelich J and Goldman J (2013) Zinc exhibits ideal sodium chloride solution. Electrochimica Acta 54(4): 1204–1209.
physiological corrosion behavior for bioabsorbable stents. Advanced 32. Orazem M and Tribollet B (2008) Electrochemical Impedance
Materials 25(18): 2577–2582. Spectroscopy. Wiley, Hoboken, NJ, USA.
13. Li HF, Xie XH, Zheng YF et al. (2015) Development of biodegradable 33. Stoilova D, Koleva V and Vassileva V (2002) Infrared study of some
Zn-1X binary alloys with nutrient alloying elements Mg, Ca and Sr. synthetic phases of malachite (Cu2(OH)2CO3)–hydrozincite (Zn5
Scientific Reports 5: 13. (OH)6(CO3)2) series. Spectrochimica Acta Part A: Molecular and
14. Shearier ER, Bowen PK, He W et al. (2016) In vitro cytotoxicity, Biomolecular Spectroscopy 58(9): 2051–2059.
adhesion, and proliferation of human vascular cells exposed to zinc. 34. Tie D, Feyerabend F, Muller WD et al. (2013) Antibacterial biodegradable
ACS Biomaterials Science & Engineering 2(4): 634–642. Mg–Ag alloys. European Cells and Materials 25(16): 284–298.
15. Beyersmann D (2002) Homeostasis and cellular functions of zinc. 35. Dambatta MS, Kurniawan D, Sudin I, Yahaya B and Hermawan
Materialwissenschaft und Werkstofftechnik 33(12): 764–769. H (2016) Influence of homogenization treatment on the

91
Downloaded by [ Universidad Nacional Autonoma de México (UNAM)] on [17/09/18]. Copyright © ICE Publishing, all rights reserved.
Surface Innovations Zn–Mg and Zn–Ag degradation
Volume 6 Issue SI1–2 mechanism under biologically relevant
conditions
Törne, Khan, Örnberg and Weissenrieder

degradation behavior of Zn–3Mg alloy in simulated body fluid 37. Törne K, Örnberg A and Weissenrieder J (2017) The influence of
solution. Proceedings of the Institution of Mechanical Engineers buffer system and biological fluids on the degradation of magnesium.
Part L: Journal of Materials: Design and Applications 230(2): Journal of Biomedical Materials Research Part B: Applied
615–619. Biomaterials 105(6): 1490–1502.
36. Aramaki K (2001) The inhibition effects of chromate-free, anion 38. Törne K, Örnberg A and Weissenrieder J (2017) Influence of strain on
inhibitors on corrosion of zinc in aerated 0.5 M NaCl. Corrosion the corrosion of magnesium alloys and zinc in physiological
Science 43(3): 591–604. environments. Acta Biomaterialia 48: 541–550.

How can you contribute?


To discuss this paper, please submit up to 500 words to the
journal office at journals@ice.org.uk. Your contribution will
be forwarded to the author(s) for a reply and, if considered
appropriate by the editor-in-chief, it will be published as a
discussion in a future issue of the journal.

ICE Science journals rely entirely on contributions from


the field of materials science and engineering. Information
about how to submit your paper online is available at
www.icevirtuallibrary.com/page/authors, where you will also
find detailed author guidelines.

92
Downloaded by [ Universidad Nacional Autonoma de México (UNAM)] on [17/09/18]. Copyright © ICE Publishing, all rights reserved.

You might also like