Sie sind auf Seite 1von 18

Acta Biomaterialia 106 (2020) 410–427

Contents lists available at ScienceDirect

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

Full length article

A biodegradable Zn-1Cu-0.1Ti alloy with antibacterial properties for


orthopedic applications
Jixing Lin a,b,1, Xian Tong a,b,1, Zimu Shi c, Dechuang Zhang b, Lishu Zhang d, Kun Wang a,
Aiping Wei a, Lufan Jin a, Jianguo Lin b, Yuncang Li e, Cuie Wen e,∗
a
Department of Material Engineering, Zhejiang Industry & Trade Vocational College, Wenzhou 325003, PR China
b
School of Materials Science and Engineering, Xiangtan University, Xiangtan 411105, PR China
c
Key Laboratory of Materials Physics, Institute of Solid State Physics, Chinese Academy of Sciences, Hefei 230031, PR China
d
Institute of Advanced Materials for Nano-Bio Applications, School of Ophthalmology & Optometry, Eye Hospital, Wenzhou Medical University, Wenzhou
325027, PR China
e
School of Engineering, RMIT University, Melbourne, Victoria 3001, Australia

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

Article history: Zinc (Zn) alloys are receiving increasing attention in the field of biodegradable implant materials due to
Received 17 September 2019 their unique combination of suitable biodegradability and good biological functionalities. However, the
Revised 19 January 2020
currently existing industrial Zn alloys are not necessarily biocompatible, nor sufficiently mechanically
Accepted 11 February 2020
strong and wear-resistant. In this study, a Zn–1Cu–0.1Ti alloy is developed with enhanced mechanical
Available online 14 February 2020
strength, corrosion wear property, biocompatibility, and antibacterial ability for biodegradable implant
Keywords: material applications. HR and HR + CR were performed on the as-cast alloy and its microstructure, me-
Antibacterial ability chanical properties, frictional and wear behaviors, corrosion resistance, in vitro cytocompatibility, and an-
Biodegradable metal tibacterial ability were systematically assessed. The microstructures of the Zn–1Cu–0.1Ti alloy after differ-
Friction and wear ent deformation conditions included a η-Zn phase, a ε -CuZn5 phase, and an intermetallic phase of TiZn16 .
In vitro biocompatibility The HR+CR sample of Zn–1Cu–0.1Ti exhibited a yield strength of 204.2 MPa, an ultimate tensile strength
Zn alloy
of 249.9 MPa, and an elongation of 75.2%; significantly higher than those of the HR alloy and the AC al-
loy. The degradation rate in Hanks’ solution was 0.029 mm/y for the AC alloy, 0.032 mm/y for the HR+CR
alloy, and 0.034 mm/y for the HR alloy. The HR Zn–1Cu–0.1Ti alloy showed the best wear resistance, fol-
lowed by the AC alloy and the alloy after HR + CR. The extract of the AC Zn–1Cu–0.1Ti alloy showed over
80% cell viability with MC3T3-E1 pre-osteoblast and MG-63 osteosarcoma cells at a concentration of ≤
25%. The as-cast Zn–1Cu–0.1Ti alloy showed good blood compatibility and antibacterial ability.

Statement of Significance

This work repots a Zn–1Cu–0.1Ti alloy with enhanced mechanical strength, corrosion wear property, bio-
compatibility, and antibacterial ability for biodegradable implant applications. Our findings showed that
Zn–1Cu–0.1Ti after hot-rolling plus cold-rolling exhibited a yield strength of 204.2 MPa, an ultimate ten-
sile strength of 249.9 MPa, an elongation of 75.2%, and a degradation rate of 0.032 mm/y in Hanks’ Solu-
tion. The hot-rolled Zn–1Cu–0.1Ti showed the best wear resistance. The extract of the as-cast alloy at a
concentration of ≤ 25% showed over 80% cell viability with MC3T3-E1 and MG-63 cells. The Zn–1Cu–0.1Ti
alloy showed good hemocompatibility and antibacterial ability.
© 2020 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.

1. Introduction

With the acceleration of the aging process in human soci-


ety, the hazards of senile osteoporosis are becoming more dan-
gerous and the risk of osteoporosis-induced skeletal fractures in

Corresponding author. these patients is increasing [1]. Osteoporosis is listed as the sec-
E-mail address: cuie.wen@rmit.edu.au (C. Wen). ond most harmful disease to human health by the World Health
1
Joint first author.

https://doi.org/10.1016/j.actbio.2020.02.017
1742-7061/© 2020 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.
J. Lin, X. Tong and Z. Shi et al. / Acta Biomaterialia 106 (2020) 410–427 411

Organization (WHO) [2]. In addition, the incidence of osteoporosis Table 1


Chemical composition of Zn–1Cu–0.1Ti alloy.
appears to be occurring at younger ages. At present, fracture fixa-
tion materials mainly consist of non-biodegradable metals such as Alloy Cu Ti Impurity Zn
stainless steels, cobalt (Co)-based alloys, and titanium (Ti) alloys, Zn–1Cu–0.1Ti 1.06 0.102 0.069 98.77
which generally remain in the body permanently or need follow-
up surgery to remove them after fracture healing, which brings un-
necessary pain and economic burden to patients and society [3,4].
tibacterial property for S. aureus. Leone et al. [23] reported that the
Thus, the use of biodegradable metal materials for bone fracture
addition of Ti to Zn alloys refined the matrix grains, thus result-
fixation is highly desirable. In the past, the research direction of
ing in enhanced mechanical properties. Simultaneously adding low
biodegradable materials mainly focused on magnesium (Mg) alloys
concentrations of Cu and Ti to Zn alloys improved their strength,
[5], iron (Fe) alloys [6], polymers [7], and bioceramics [8]. However,
hardness, and creep resistance, the latter also a highly desired ma-
due to the low ultimate tensile strength of polymer materials and
terial property for implant application [22]. However, most of the
the low toughness of bioceramic parts [7,8], these materials do not
existing Zn alloys still have inadequate material properties (e.g.,
meet the mechanical property requirement for bone fracture fixa-
mechanical, corrosion, biological, and antibacterial properties) for
tion. On the contrary, metallic materials possess good mechanical
implant material applications.
properties such as high strength and toughness, which could be
Although patient antibiotic prophylaxis and sterilization of im-
utilized for bone fixation; therefore, the development of biodegrad-
plants are generally performed before implantation, the failure rate
able metal materials for bone tissue engineering applications has
of implantation due to bacterial infection at the initial stage is still
become a hot research topic [9,10]. Research on biodegradable Mg,
as high as 4∼6% because once a bacterial infection occurs at the
Fe, and zinc (Zn)-based alloys is receiving increasing attention and
site of implantation, it is challenging to treat [27]. Postoperative
has achieved substantial advances in terms of their mechanical
infections with implants are considered a significant challenge in
properties and biocompatibility in recent years. Nonetheless, both
implant surgery. Therefore, an implant material with intrinsic an-
Mg and Fe alloys have certain disadvantages; for instance, most
tibacterial ability is highly desirable. Ren et al. [28] systematically
Mg alloys exhibit high degradation rates and Fe alloys degrade too
investigated a variety of Cu-containing medical metals such as
slowly in the physiological environment [11]. However, it is worth
stainless steels, Ti alloys, and Co-based alloys, and demonstrated
noting that MgYREZr alloy devices have already been successfully
the good antibacterial ability of these materials stemming from
commercialized such as Bioabsorbable interference screws (MAG-
the durable and broad-spectrum antibacterial characteristics of
NEZIX®, and Magmaris by BIOTRONIK), but Zn-based biodegrad-
Cu ions. Therefore, Cu-containing Zn alloys can be expected to be
able materials are still in the development stage [12]. On the other
promising implant materials with intrinsic antibacterial ability.
hand, Zn and its alloys have moderate corrosion rates between
After implantation, there is always a certain degree of micro-
those of Mg-based and Fe-based alloys.
movement at the bone-implant interface, resulting in wear de-
Zn is an essential trace element in the human body [13]. As a
bris which may cause adverse tissue responses. Moreover, interface
component of more than 300 kinds of enzymes in the body, Zn
damage to implants may also occur due to wear and chemical cor-
participates in a variety of biological activities such as maintain-
rosion in the physiological environment. Therefore, the corrosion
ing human immune function, growth, and development, promot-
and micro-abrasive wear behaviors of implant materials should be
ing insulin secretion, and enhancing memory [14–16]. More im-
investigated. The corrosion wear performances of stainless steels
portantly, Zn can also improve bone tissue development and have
[29], Ti alloys [30,31], Co–Cr–Mo alloy [32], and Mg alloy [33] have
an osteoinductive effect on bone tissue [17]. Hence, Zn and its al-
been reported in recent years. However, the corrosion wear behav-
loys can provide potential biological functions as implant materi-
ior of Zn alloys is still rarely reported.
als for medical applications such as bone fixation and cardiovas-
In this study, we developed a new Zn–1Cu–0.1Ti alloy for
cular stents. However, pure Zn has extremely low yield strength
potential biodegradable implant applications. Plastic deformation
(∼29.3 MPa) and elongation (∼1.2%) in its as-cast condition due
processes including hot-rolling and hot-rolling followed by cold-
to its hexagonal close-packed (hcp) crystal structure [18]. To date,
rolling were performed on the as-cast alloy, and its microstruc-
approaches for effectively improving the mechanical properties of
ture, mechanical properties, corrosion resistance, friction and wear
pure Zn include plastic deformation and the addition of alloying el-
behaviors in both Hanks’ solution and a dry-sliding friction envi-
ements such as germanium (Ge) [18], Mg [19,20], calcium (Ca) [19],
ronment, antibacterial ability, and in vitro cytotoxicity after various
strontium (Sr) [19], copper (Cu) [21,22], and Ti [23]. Venezuela
plastic deformation processes were systematically investigated.
and Dargusch [24] comprehensively reviewed the influence of al-
loying and fabrication techniques on the mechanical properties,
biodegradability and biocompatibility of Zn alloys, and suggested 2. Experimental procedure
that the wrought Zn-xMg alloy (x < 1.5%) exhibits the best com-
bination of mechanical properties, biodegradability, and biocom- 2.1. Material preparation
patibility. Li et al. [25] reviewed the mechanical aspects of indus-
trial and biomedical Zn and its alloys and recommended that more A Zn alloy with a nominal composition of Zn–1Cu–0.1Ti (wt.%
mechanical property tests are needed to evaluate the mechanical hereafter) was prepared by casting using pure Zn (purity 99.995%,
properties of biodegradable Zn-based metals such as creeping re- Huludao Zinc Industry Co. Ltd., China), pure Cu (99.99%, Jiangxi
sistance, corrosion fatigue, and stability of mechanical properties Copper Co. Ltd., China), and a Zn–2Ti master alloy (Hunan High
during storage and service. Hernández-Escobar et al. [26] reviewed Broad New Material Co. Ltd., China) as starting materials. The raw
the most recent progress on the development of absorbable Zn- materials were melted in a graphite crucible at 550–600 °C un-
based alloys for biomedical implant applications and suggested to der the protection of argon (Ar) gas and then cast into a square
re-evaluate the mechanical requirements and define the standard mold which was preheated to 250 °C. The chemical composition
specifications of the absorbable metal implant, which are indepen- of the as-cast Zn–1Cu–0.1Ti alloy is listed in Table 1, as mea-
dent with those of the permanent implants. Tang et al. [21] inves- sured by X-ray fluorescence (XRF; S4 Pioneer, Bruker, Germany).
tigated a series of biodegradable Zn–x Cu (x = 1, 2, 3, 4 wt.%) alloys An as-cast ingot with dimensions of 85 mm × 85 mm × 20 mm
and reported a good elongation of ∼50.6% for a Zn–4Cu alloy, along was annealed at 340 °C for 10 h for homogenization, followed by
with acceptable cytotoxicity for human endothelial cells and an- cooling to room temperature in air. Plates with dimensions of
412 J. Lin, X. Tong and Z. Shi et al. / Acta Biomaterialia 106 (2020) 410–427

60 mm × 20 mm × 10 mm were cut from the alloy ingot for hot- 2.5. Immersion testing
rolling and hot-rolling plus cold rolling using electrical discharge
machining (EDM). The alloy plates were heated to 250 °C and held The degradation rates of the AC, HR, and HR+CR samples of Zn–
for 5 min before each hot-rolling process. The as-cast Zn–1Cu–0.1Ti 1Cu–0.1Ti were measured by immersion tests. disk samples with
alloy plates were divided into two groups for plastic deformation: dimensions of diameter 8 mm and thickness 2 mm were cut by
the first group of plates was hot-rolled to a total reduction of 80% EDM and polished up to 20 0 0 grit. Immersion tests were carried
in thickness at 0.5 mm reduction per pass; and the second group out by soaking the samples in Hanks’ solution for 14 and 30 d at
of plates was hot-rolled to a total reduction of 40% in thickness at 37 °C, where the ratio of specimen surface area to solution volume
0.5 mm reduction per pass and then air-cooled to room tempera- was 1 cm2 per 20 ml. The morphology and composition of the cor-
ture, followed by cold-rolling to a total reduction of 40% in thick- rosion products on the sample surfaces after immersion tests were
ness at 0.2 mm reduction per pass with room temperature of 25 °C. characterized by SEM and EDS. Ultrasonic cleaning using a mixed
Both groups of plate samples exhibited a final thickness of 2 mm. solution of CrO3 (200 g/L) and AgNO3 (10 g/L) was used to remove
the corrosion products after immersion tests. The weights of the
2.2. Microstructure characterization samples before and after immersion tests were measured by an
electronic balance with an accuracy of 0.1 mg (Sartorius BSA124S-
The microstructures of the as-cast, hot-rolled, and hot-rolled CW, Germany) and the degradation rate was calculated based on
plus cold-rolled (denoted AC, HR, and HR+CR, respectively here- ASTM G31-72 [37]. Measurements were performed at least in trip-
after) Zn–1Cu–0.1Ti alloy samples were characterized by optical licate. The phases of the corrosion products of HR Zn–1Cu–0.1Ti
microscopy (OM; DM2500C, Leica, Germany). All samples were alloy after immersion in Hanks’ solution for 30 d were analyzed
polished and etched with a 0.5 vol.% HNO3 ethanol solution. The by a micro-area XRD diffractometer (micro-XRD; D/Max Rapid IIR,
microstructures of the HR and HR+CR samples were observed in Rigaku, Japan) at a voltage of 40 kV and a current of 150 mA with
the rolling direction. X-ray diffraction analysis (XRD; D8 ADVANCE, Cu–Kα radiation at a step size of 0.1 s−1 . The XRD data were over a
Bruker, Germany) with Cu–Kα radiation was performed to identify 2θ range of 0–90 °. Micro-Fourier transform infrared spectroscopy
the phases of the alloy samples, and the scanning spectra were (micro-FTIR; Nicolet 6700, Thermo Fisher, USA) with a wave range
recorded over a 2θ range from 10 ° to 90 ° at a step size of of 40 0 0–650 cm−1 at a resolution of 4 cm−1 was performed to
0.02 °s−1 . Scanning electron microscopy (SEM; Pro-X FEI, Phenom, identify the functional groups in the corrosion products. X-ray pho-
Netherlands) along with energy dispersive spectrometry (EDS; X- toelectron spectroscopy (XPS; Thermo Fisher Scientific, USA) with
Max, Oxford, England) were carried out to evaluate the microstruc- an Al–Kα source at a step size of 1 eV was used to analyze the
tures and chemical compositions of the phases of the alloy samples chemistry of the corrosion products. High-resolution narrow scans
at 15 keV. at a step size of 0.1 eV were performed to determine the binding
states of Zn 2p, O 1 s, Ca 2p, and C 1 s of the corrosion products.
2.3. Mechanical testing
2.6. Friction and wear testing
The tensile properties of the AC, HR, and HR+CR samples of
Zn–1Cu–0.1Ti were carried out using a universal testing machine The friction and wear behaviors of the AC, HR, and HR+CR sam-
(Instron-3369, USA) at a strain rate of 1 mm/min at room tempera- ples of Zn–1Cu–0.1Ti were investigated using a high-speed recip-
ture. The dimensions of the tensile samples were prepared accord- rocating friction tester (HSR-2M; Lanzhou Zhongke Kaihua Tech-
ing to ASTM E8/E8M-08 [34] and the samples after deformation nology Development Co., China). For comparison, as-cast pure Zn
were cut by EDM along the rolling direction for microstructural ob- (hereafter denoted AC pure Zn) was also tested under the same
servation. The fracture surfaces of the samples after tensile testing conditions. The tests were carried out under a wear load of 200 gf,
were observed by SEM. The hardness of the AC, HR, and HR+CR a friction stroke length of 5 mm, a friction frequency of 1 Hz, a test-
alloy samples was measured using a Vickers microhardness tester ing time of 10 min, and a friction pair of ZrO2 with a diameter of
(MicroMet 60 0 0, Buehler, USA) with a load of 10 0 gf and a holding 3 mm. Disk samples having 15 mm diameter and 10 mm thickness
time of 15 s. were cut by EDM and mounted with epoxy resin, and their sur-
faces were polished to a roughness (Ra) of 0.5 μm using diamond
2.4. Electrochemical testing paste before wear testing. Both sample surfaces and the friction
pair were cleaned using ethanol before friction and wear testing.
The corrosion behaviors of the AC, HR, and HR+CR samples Both dry-sliding wear tests and corrosion wear tests in a liquid lu-
of Zn–1Cu–0.1Ti in Hanks’ solution [35] were investigated by po- brication environment were performed on the wear-testing sam-
tentiodynamic polarization testing and electrochemical impedance ples. The corrosion wear tests were carried out by immersing the
spectroscopy (EIS) testing with a pH of 7.40 at 37 °C using an samples in a sink containing 50 ml Hanks’ solution at a constant
electrochemical workstation with a three-electrode electrochemi- temperature of 37 °C, and the dry-sliding friction wear tests were
cal cell (ParStat 2273, Princeton Applied Research, USA). The sam- performed at room temperature. The weights of the samples before
ples were prepared by grinding with SiC paper up to 20 0 0 grit and after the friction and wear tests were measured using an elec-
and polishing to 0.5 μm with diamond paste. During electrochem- tronic balance to determine the weight loss. The surface morphol-
ical testing, samples with an exposed area of 0.5 cm2 were set as ogy and roughness after friction and wear tests were characterized
the working electrode, a platinum plate electrode as the counter by SEM and a 3D-contour testing software.
electrode, and a saturated calomel electrode (SCE) as the reference
electrode. The potential polarization tests were performed over a 2.7. Surface contact angle testing
scan range from −800 mV to 500 mV relative to the open circuit
potential (OCP) at a potential scan rate of 1 mV s–1 . EIS measure- The surface contact angle of the AC, HR, and HR+CR samples of
ments were carried out with a disturbance frequency ranging from Zn–1Cu–0.1Ti and AC pure Zn was measured using a contact angle
100 kHz to 0.01 Hz and a perturbation signal-amplitude of 10 mV goniometer (OCA15EC, Data Physics, Germany) at room tempera-
at OCP. The corrosion rate of the samples was calculated based on ture. Alloy samples with dimensions of diameter 8 mm and thick-
ASTM G102-89 [36]. All procedures were done at least in triplicate ness 2 mm were ground with 20 0 0 grit sandpaper. A 4 μl droplet
samples. of deionized water dyed with ink was used as the testing medium.
J. Lin, X. Tong and Z. Shi et al. / Acta Biomaterialia 106 (2020) 410–427 413

Three points were randomly selected on each sample for testing, according to ISO 10993-5: 2009 [39]. The MC3T3-E1 or MG-63
and the average value was taken as the result of surface contact cells were seeded in a 96-well plate with 5 × 103 cells per well
angle. and cultured for 24 h. The original medium was replaced with the
diluted extracts of different concentrations of 100%, 25%, and 12.5%
2.8. Hemolysis testing and then incubated for 1, 3, and 5 d for evaluation of cell viability.
Both the cell assays were done at least in triplicate.
Hemolysis tests of the AC, HR, and HR+CR samples of Zn–1Cu– After incubation for the corresponding time for both direct and
0.1Ti and AC pure Zn were carried out according to ISO 10993- indirect cell assays, 10 μl CCK-8 solution (Dojindo, Japan) was
4: 2002 [38]. Extracts of the alloy samples were prepared using added to each well and then incubated for 2 h. Optical density was
0.9% normal saline solution at an extraction ratio of 5 g/10 ml. The measured at the wavelength of 450 nm using a microplate reader
hemolysis test procedures were carried out under a strict protocol (Gen5, Biotek, USA). The cell viability of the MC3T3-E1 and MG-63
at Wenzhou Medical University under the guidance of its Ethics cells of the pure Zn and Zn–1Cu–0.1Ti was calculated according to
Committee, which gave ethics approval beforehand. Fresh blood the following formula [39]:
from healthy humans (12 ml) had 0.4 ml of 20 g/l potassium ox-
alate reagent added and the solution washed with 30 ml of nor- Cell viability (% ) = ( (ODt − ODblank )/(ODnc − ODblank ) )×100%
mal saline three times, then centrifuged to obtain red blood cells. (2)
Finally, normal saline was used to configure a 16% red blood cell
where ODt is the absorbance value of the experimental group;
suspension. The metal extract (4 ml) was incubated for testing, and
ODnc is the absorbance value of the negative control group; and
the deionized water and 0.9% sodium chloride solution were used
ODblank is the absorbance value of the α -MEM medium.
as the positive and negative controls, respectively. All centrifuge
tubes were placed in a water bath at 37 °C for 30 min, then 0.2 ml
2.10. Antibacterial evaluation
of the 16% red blood cell suspension was added to each centrifuge
tube and all tubes were continually cultured at 37 °C for 60 min.
The antibacterial properties of as-cast Zn–1Cu–0.1Ti and pure
The samples were centrifuged at 10 0 0 r/min for 5 min at 25 °C un-
Zn were evaluated by the inhibition zone diameter (IZD) method
til erythrocytes precipitated, then 200 μl of the supernatant was
according to DIN EN ISO 20645-2004 [40]. Staphylococcus aureus
pipetted into a 96-well plate. The absorbance values were mea-
(S. aureus) (CMCC26003) was placed into 5 ml of medium and
sured by a microplate photometer (Multiskan MK3, Thermo Scien-
shaken for 24 h in a constant temperature shaker, then the bac-
tific, USA) at a wavelength of 545 nm, and the hemolysis percent-
terial solution was diluted to a concentration of 1 × 108 CFU/ml
age was calculated according to the following formula:
with phosphate-buffered saline (PBS) buffer. Then Luria-Bertani
Hemolysis percentage (LB) broth powder (10 g/l peptone, 5 g/l yeast extract powder,
  10 g/l sodium chloride) was added to 950 ml distilled water, then
= absorbanceexperimental group − absorbancenegative group /
  adjusted to pH 7.0∼7.2 with 0.1 mol/l NaOH solution after en-
absorbancepositive group − absorbancenegative group ×100% (1) tirely dissolving and stirring all contents, and then made up to
a volume of 10 0 0 ml with distilled water. The medium solution
2.9. Cytotoxicity evaluation and the pure Zn and Zn–1Cu–0.1Ti alloy disk samples (diame-
ter 8 mm × thickness 2 mm) were autoclaved at 121 °C for 40 min
The in vitro tests were performed by both indirect and di- and sterilized by UV sterilization on a sterile bench for 30 min.
rect contact assays using MC3T3-E1 pre-osteoblast cells (mouse Agar powder (1.5 g per 100 ml of the medium) was added to the
cell lines, NIH/3T3, GDC030) and MG-63 osteosarcoma cells (hu- medium, and then autoclaved for 30 min after heating and dissolv-
man cell lines, MG-63, GDC074) (purchased from cell bank of Chi- ing. The medium solution (20 ml) was poured into a culture dish at
nese Academy of Sciences, Shanghai, China). Cells were cultured a temperature of 45 °C and UV-sterilized for 30 min to prepare an
in alpha-minimum essential medium (α -MEM; 0.1 mM, Invitrogen, agar medium plate. The alloy sample was placed in the center of
Carlsbad, CA) supplemented with 10% fetal bovine serum (FBS; the plate and 500 μl of the bacterial suspension was evenly spread
Lonsera; Cat Number: S711-001S) and antibiotics (100 units/ml on the surface of the agar medium with a pipette. At least five
penicillin and 100 μg/ml streptomycin; Aladdin; Cat Number: times duplicates were measured for statistical analysis. The IZD
PH1513) in a humidified atmosphere of 5% CO2 at 37 °C for 72 h. values were observed and measured after incubation at 37 °C for
disk samples of as-cast Zn–1Cu–0.1Ti and pure Zn with dimensions 24 h, and the antibacterial properties were calculated according to
of diameter 8 mm and thickness 2 mm were prepared for the cyto- the formula given by [40]:
toxicity assessment. The disk samples were ultrasonically cleaned
in analytically pure ethanol with a concentration of 99.7% and ster- H = (D − d )/2 (3)
ilized in ultraviolet (UV) radiation before testing. where H is the inhibition zone in mm, D is the total diameter of
For the direct cell assay, MC3T3-E1 cell droplets the specimen and inhibition zone in mm, and d is the diameter of
(3 × 104 cells/well) were seeded on the surfaces of the disk the specimen in mm. When H is equal to or larger than 1 mm and
samples and cultured at 37 °C in 95% relative humidity with 5% there is no growth of bacteria, the antibacterial property is good;
CO2 for 24 h and 48 h, respectively. To testing the reducibility of when H is equal to 0 mm and there are regions with some bacteria,
the metal disk, blank absorbance was measured at 450 nm in the the antibacterial property is limited; and when H is equal to 0 mm
culture medium and CCK-8 solution without cells but with metal and there are regions with many bacteria, there is no antibacterial
samples in the same cultural environment. If the absorbance is property.
minimal, it means that CCK-8 solution has not been reduced and
can be directly added for testing. Before the CCK-8 assay, the origi- 2.11. Statistical analysis
nal culture medium was discarded to avoid the interference of the
reduction effect from metal samples. The cells were washed twice In this study, the experimental results are expressed as means
with fresh culture medium, and then finally replaced with culture ± standard deviations. One-way analysis of variance (V17.0, SPSS,
medium. For the indirect cell assay, the extracts were prepared USA) was applied to determine the statistical significance of the
by immersing the disk samples of as-cast Zn–1Cu–0.1Ti and pure differences observed between groups. P < 0.05 was accepted as
Zn in the culture medium for 72 h at a ratio of 1.25 cm2 ml−1 statistically significant.
414 J. Lin, X. Tong and Z. Shi et al. / Acta Biomaterialia 106 (2020) 410–427

Table 2
EDS analysis results of different spots in Zn–1Cu–
0.1Ti alloy samples.

Spot no. Zn (at.%) Cu (at.%) Ti (at.%)

Spot 1 90.4 ± 0.3 0.4 ± 0.2 9.2 ± 0.2


Spot 2 92.3 ± 1.4 0.4 ± 0.1 7.3 ± 1.3
Spot 3 97.0 ± 0.4 0.9 ± 0.1 2.1 ± 0.3
Spot 4 98.6 ± 0.1 1.3 ± 0.0 0.1 ± 0.1
Spot 5 88.6 ± 0.1 7.7 ± 0.1 3.7 ± 0.0
Spot 6 87.0 ± 0.4 13.0 ± 0.4 0.00

the η-Zn phase (as marked in Fig. 2(b)), indicating that incom-
plete dynamic recrystallization and grain growth occurred during
hot rolling. The HR+CR sample shows clear processing flow lines
and black ε -CuZn5 particles distributed in the matrix along the
rolling direction (Fig. 2(c)). In addition, the ε -CuZn5 phase is re-
fined compared to that of the HR sample.
Fig. 1. XRD patterns of AC, HR, and HR + CR samples of Zn–1Cu–0.1Ti.
Fig. 3 shows SEM images of the AC, HR, and HR+CR samples
of Zn–1Cu–0.1Ti. The EDS analysis results for the different micro
spots in Fig. 3 are shown in Table 2. From the SEM images and
the corresponding EDS data, the short rod-like and granular phases
3. Results corresponding to spots 1, 2, and 3 in Fig. 3(a) are mainly com-
posed of Zn, Ti, and a small amount of Cu. The amount of Ti in
3.1. Microstructure of Zn–1Cu–0.1Ti the three spots varies significantly and its relative atomic content
is 9.2 ± 3.2% for spot 1, 7.3 ± 1.3% for spot 2, and 2.1 ± 0.3% for spot
Fig. 1 shows the XRD patterns of the AC, HR, and HR+CR sam- 3, respectively. Because the particle size of the phase in spot 3 is
ples of Zn–1Cu–0.1Ti. It can be seen that all of the samples con- very small, a portion of the electron beam was irradiated onto the
tain an η-Zn phase (the matrix), an intermetallic compound TiZn16 Zn matrix during EDS analysis, resulting in a higher Zn content in
phase, and an ε -CuZn5 phase. Owing to the small additions of Cu this spot. According to the XRD patterns and the Zn–Ti phase dia-
and Ti in the Zn–1Cu–0.1Ti alloy, the intensity of the diffraction gram [41,42], it can be suggested that the short rod-like and gran-
peaks of the ε -CuZn5 phase and the TiZn16 phase in the XRD pat- ular Ti-rich phase is the TiZn16 phase. Based on the Zn–Ti phase
terns of the AC and HR Zn–1Cu–0.1Ti samples is weak, and all the diagram, the solid solubility of Ti in a Zn matrix is very small (only
diffraction peaks of the η-Zn matrix phase show similar intensities. 0.0 07∼0.015 wt.% at 30 0 °C [42]). When the Ti content is higher
As for the HR+CR sample, the η-Zn phase shows significantly en- than 0.21 wt.% and the temperature reaches about 418.6 °C, a eutec-
hanced intensity for the (004) peak but decreased intensity for the tic reaction occurs in the Zn–1Cu–0.1Ti alloy and the TiZn16 phase
other peaks; and the ε -CuZn5 phase shows increased peak inten- forms with an excess of Ti and Zn, which is distributed on the
sities and its main peak shows an intensity similar to that of the grain boundary and inside of the grains. At the same time, due
main peak (101) of the η-Zn phase. to the faster solidification of the Zn–1Cu–0.1Ti alloy, the Ti un-
Fig. 2 shows the optical microstructures of the AC, HR, and derwent an unequilibrium eutectic reaction, resulting in entirely
HR+CR samples of Zn–1Cu–0.1Ti. It can be seen that the mi- different Ti contents in the different precipitation phase, concen-
crostructure of the AC sample consists mainly of a white irregu- trated in local areas such as eutectic branches and forming severe
lar η-Zn matrix phase with an average size of 53.9 ± 7.6 μm and a segregation. Spot 4 is a Zn-rich phase that contains small amounts
short gray rod or granular eutectic intermetallic compound TiZn16 of Cu and Ti, and can be inferred to be an η matrix phase. The
phase (Fig. 2(a)). Most of the eutectic TiZn16 precipitates are dis- rod-shaped phase shown in spot 5 (Fig. 3(b)) is mainly composed
tributed along the grain boundaries, with a small amount of the of Zn, Cu, and Ti, and the relative atomic contents of Cu and Ti
TiZn16 phase distributed in the grain matrix of the η-Zn phase. As are 7.7 ± 0.1% and 3.7 ± 0.0%, respectively. The atomic contents of
for the microstructure of the HR sample, the η-Zn matrix phase these two elements are similar to those reported in the ternary in-
and eutectic TiZn16 precipitates are elongated along the rolling di- termetallic compound of Cu2 TiZn22 by Zermout et al. [43]. It can
rection, and the TiZn16 phase shows the morphology of fibers. The be deduced that the phase in spot 5 is Cu2 TiZn22 . Since the con-
size of the η-Zn phase is increased as compared to the AC sam- centration of the Cu2 TiZn22 phase is small, there are no significant
ple, and the average size reaches 84.4 ± 12.5 μm. Also, many cellu- diffraction peaks observable in the XRD spectra. The finer granular
lar crystal grains with inconspicuous grain boundaries appear in phase of spot 6 (in Fig. 3(c)) has only Zn and Cu elements, and the

Fig. 2. Microstructures of Zn–1Cu–0.1Ti alloy samples: (a) AC; (b) HR; and (c) HR + CR.
J. Lin, X. Tong and Z. Shi et al. / Acta Biomaterialia 106 (2020) 410–427 415

Fig. 3. SEM images of Zn–1Cu–0.1Ti alloy samples: (a) AC; (b) HR; and (c) HR + CR.

atomic fraction of Cu was 13.0 ± 0.4%. These atomic ratios are close spectively for the AC sample; 205.7 ± 5.5 MPa, 175.4 ± 3.8 MPa,
to those of the CuZn5 phase. Based on the results from XRD, EDS, 39.2 ± 1.4%, and 70.6 ± 1.8 HV, respectively for the HR sample; and
and the Zn–Cu phase diagram [44], it can be concluded that this 249.9 ± 3.8 MPa, 204.2 ± 4.3 MPa, 75.2 ± 1.9%, and 56.4 ± 0.8 HV,
intermetallic phase is CuZn5 . respectively for the HR+CR sample (Fig. 4(b)). The σ uts and ε
of the HR+CR sample increased by 21.5% and 91.8%, respectively
3.2. Mechanical properties of Zn–1Cu–0.1Ti as compared to those of the HR sample. It can be seen that the
HR+CR sample exhibits extraordinarily high ductility with an
Fig. 4 shows the tensile stress-strain curves, yield strength ε = 75.2 ± 1.9%. The SEM images of the fracture surfaces of the AC,
(σ ys ), ultimate tensile strength (σ uts ), elongation (ε ), hardness, HR, and HR+CR samples are shown in Fig. 4(c)–(e). The fracture
and fracture surfaces of the AC, HR, and HR+CR samples of the surface of the AC sample shows a morphology of cleavage fractures
Zn–1Cu–0.1Ti alloy. The σ uts , σ ys , ε , and hardness values were with river patterns and cleavage steps, and some small dimples
92.4 ± 4.4 MPa, 86.1 ± 2.6 MPa, 1.4 ± 0.8%, and 72.6 ± 0.6 HV, re- in localized areas (Fig. 4(c)). It can be inferred that the AC sample

Fig. 4. Mechanical properties and fracture surfaces of Zn–1Cu–0.1Ti alloys (a) tensile stress-strain curves; (b) tensile yield strength, ultimate strength, elongation, and hard-
ness; (c) SEM image of AC fracture surface; (d) SEM image of HR fracture surface; and (e) SEM image of HR+CR fracture surface.
416 J. Lin, X. Tong and Z. Shi et al. / Acta Biomaterialia 106 (2020) 410–427

exhibits a mixed mode of fracture with brittle cleavages and a 3.3. Corrosion behavior of Zn–1Cu–0.1Ti
few ductile dimples. In the HR and HR+CR samples, the fracture
surfaces are homogeneously distributed with round, equiaxed Fig. 5 shows the corrosion behaviors of the AC, HR, and
dimples of different sizes and depths. The fracture mode of these HR+CR samples of Zn–1Cu–0.1Ti in Hanks’ solution at 37 °C.
samples is typical ductile microvoid coalescence, indicating that The OCP and potentiodynamic polarization curves of the sam-
thermomechanical processes can significantly improve the duc- ples are shown in Fig. 5(a) and (b), respectively. The corre-
tility of the Zn–1Cu–0.1Ti alloy. Furthermore, the dimples on the sponding corrosion performance indicators obtained by Tafel re-
fracture surfaces of the HR+CR sample are denser, deeper, and gion extrapolation are listed in Table 3. The OCP curves of the
larger compared to those of the HR sample and the area occupied AC, HR, and HR+CR samples of Zn–1Cu–0.1Ti reach a steady
by the dimples is larger than that of HR sample, indicating a better fluctuation state after immersion for 1200s, and the OCP val-
plastic deformation ability. ues of the samples in descending order are AC > HR+CR

Fig. 5. Corrosion behaviors of AC, HR, and HR + CR samples of Zn–1Cu–0.1Ti in Hanks’ solution: (a) OCP curves; (b) potentiodynamic polarization curves; (c) Nyquist plots;
(d) Bode impedance modulus diagrams; (e) Bode phase angle diagrams; and (f) degradation rate after immersion for 14 d and 30 d.
J. Lin, X. Tong and Z. Shi et al. / Acta Biomaterialia 106 (2020) 410–427 417

Table 3
Electrochemical performance parameters of AC, HR, and HR+CR samples of Zn–1Cu–0.1Ti in Hanks’ solution.

Sample Corrosion potential (Ecorr ), V vs. SCE Corrosion current density (Icorr ) (μA/cm2 ) Corrosion rate (Vcorr ) (μm/y)

AC −1.025 ± 0.264 21.5 ± 0.4 315 ± 6


HR −1.123 ± 0.185 111.2 ± 0.9 1628 ± 13
HR + CR −1.100 ± 0.201 67.7 ± 0.5 991 ± 7

Table 4
Fitted results of EIS spectra for AC, HR, and HR + CR samples of Zn–1Cu–0.1Ti in Hanks’ solution.

Samples R s ( ) R f ( ) CPEf (μF) n1 Rct () CPEd1 (μF) n2 Vcorr (μm/y) Chi square

AC 1.6 ± 0.2 2667.2 ± 3.1 19.8 ± 0.4 0.68 ± 0.10 1980.4 ± 3.3 1166.7 ± 3.9 0.63 ± 0.14 263 ± 1 5.33 × 10−3
HR 0.9 ± 0.2 467.4 ± 2.3 80.0 ± 0.7 0.54 ± 0.07 1097.2 ± 5.6 1098.7 ± 3.9 0.73 ± 0.05 475 ± 5 5.15 × 10−3
HR+CR 1.4 ± 0.3 1629.0 ± 3.8 13.1 ± 0.2 0.82 ± 0.08 1426.1 ± 2.5 219.8 ± 0.7 0.77 ± 0.08 365 ± 1 8.29 × 10−3

> HR. From the potentiodynamic polarization tests, the corro- order is AC > HR + CR > HR and the HR sample showed a signif-
sion potential, corrosion current density, and corrosion rate are icantly lower value than those of both AC and HR + CR samples.
−1.025 ± 0.264 V, 21.5 ± 0.4 μA/cm2 , and 315 ± 6 μm/y for the AC Meanwhile, the phase angles of all the samples were less than or
sample; −1.123 ± 0.185 V, 111.2 ± 0.9 μA/cm2 , and 1628 ± 13 μm/y equal to 60° in this frequency region; while the phase angles of
for the HR sample; and −1.100 ± 0.201 V, 67.7 ± 0.5 μA/cm2 , and AC and HR+CR samples were higher than 60° in high-frequency
991 ± 7 μm/y for the HR+CR sample, respectively. The AC sample region, which confirmed the presence of a stable, compact film
shows the lowest current density and corrosion rate, and high- of corrosion products of the sample surface [47]. The Vcorr of the
est self-corrosion potential, and the HR sample shows the highest samples determined by the EIS spectra via Stern-Geary method is
current density and corrosion rate, and the lowest self-corrosion 263 ± 1 μm/y for the AC sample, 475 ± 5 μm/y for the HR sample,
potential. The self-corrosion potential results are consistent with and 365 ± 1 μm/y for the HR+CR sample, which is similar to the
the OCP results. Furthermore, there is a narrow passivation re- trend of Vcorr measured by the polarization test. Based on the re-
gion on the anodic polarization curve for all the samples, with a sult of EIS spectra, the corrosion resistance of the samples in de-
breakdown potential of −0.739 V, −0.826 V, and −0.821 V for AC, scending order is AC > HR + CR > HR, which is consistent with
HR, and HR+CR, respectively. In addition, the blunt current den- the potentiodynamic polarization results.
sity in ascending order is 0.0587 mA/cm2 for AC  0.1557 mA/cm2 Fig. 5(f) shows the average degradation rate of the AC, HR, and
for HR + CR < 0.4140 mA/cm2 for HR. The current density of HR + CR samples of Zn–1Cu–0.1Ti after immersion in Hanks’ so-
the HR sample is significantly higher than those of the AC and lution for 14 d and 30 d It can be seen that the degradation rate
HR + CR samples. The blunt current density and the magnitude after immersion for 14 d is 0.094 ± 0.010 mm/y for the AC sam-
of the breakdown potential are related to the stability of the passi- ple, 0.149 ± 0.012 mm/y for the HR sample, and 0.113 ± 0.015 mm/y
vation films on the metals. A lower blunt current density and more for the HR+CR sample, and the degradation rate after im-
positive breakdown potential in the passivation state correspond to mersion for 30 d was 0.029 ± 0.022 mm/y for the AC sample,
a lower dissolution rate of the metal and more stable passivation 0.034 ± 0.018 mm/y for the HR sample, and 0.032 ± 0.011 mm/y
film. According to the magnitude of the breakdown potential and for the HR+CR sample. The degradation rate of all three sam-
the blunt current density, the passivation film of the AC sample ples significantly decreased with increasing immersion time. Also,
shows the highest stability among the three samples. Therefore, it the AC sample shows the lowest degradation rate compared with
can be concluded that the corrosion resistance of the samples in the HR+CR and HR samples for immersion for both 14 d and
descending order is AC > HR + CR > HR. 30 d In addition, the HR sample shows the lowest corrosion re-
Fig. 5(c) shows the Nyquist plots of the AC, HR, and HR+CR sistance with the highest degradation rate. The degradation rate
samples of Zn–1Cu–0.1Ti. Table 4 lists the fitted results of EIS of Zn–1Cu–0.1Ti after immersion in Hanks’ solution for 30 d is
spectra for AC, HR, and HR+CR samples of Zn–1Cu–0.1Ti. It can similar to that of the extruded Zn–1Cu alloy in c-SBF solution
be seen that there are two different capacitive arcs in both the (0.033 mm/y) reported by Tang et al. [21]. Based on the results
low-frequency region and the high-frequency region in the studied from polarization, EIS, and immersion tests, the corrosion resis-
samples. The semicircular capacitance arc in the high-frequency tance performance of the Zn–1Cu–0.1Ti samples in descending or-
region is related to the charge transfer in the corrosion process, der is AC > HR + CR > HR.
and the semicircular capacitance arc in the low-frequency region Fig. 6 shows the SEM images of the surface morphologies and
is associated with the diffusion process of the passivation layer cross-sections of the AC, HR, and HR+CR samples of Zn–1Cu–0.1Ti
[45,46]. The AC sample shows the largest capacitive arc radius and after immersion in Hanks’ solution for 14 d and 30 d As can be
the HR sample shows the smallest. An equivalent circuit consisting seen, all the sample surfaces had granular or massive corrosion
of the resistance (R) and constant phase element (CPE) is shown products deposited after the immersion tests. After immersion in
in Fig. 5(c), where Rct and CPEd1 are the electric double layer re- Hanks’ solution for 14 d and 30 d, the overall thickness of the
sistance and capacitance between the electrode surface and solu- corrosion products on the sample surfaces in descending order is
tion, Rf and CPEf are the resistance and capacitance of the corro- HR > HR + CR > AC; similarly, the HR sample surface showed the
sion products on the sample surface, respectively, and Rs is the greatest amount of corrosion products with an overall thickness of
solution resistance. As shown in Table 4, the impedance of the 13.68 ± 0.39 μm and 31.96 ± 3.20 μm after immersion for 14 d and
samples in descending order is AC > HR+CR > HR and a higher 30 d, respectively. Also, the amount and overall thickness of corro-
impedance value indicates a stronger corrosion resistance. Fig. 5(d) sion products on the sample surfaces after immersion for 30 d in-
and Fig. 5(e) show the Bode impedance modulus and phase angle creased as compared to immersion for 14 d. Also, the amount and
diagrams of the AC, HR, and HR+CR samples of Zn–1Cu–0.1Ti. In overall thickness of corrosion products on the sample surfaces af-
the low-frequency region and intermediate-frequency region (0.01 ter immersion for 30 d increased compared to immersion for 14 d.
∼ 100 Hz), the impedance modulus of the samples in descending It is worth noting that there are no visible cracks or partial spalling
418 J. Lin, X. Tong and Z. Shi et al. / Acta Biomaterialia 106 (2020) 410–427

Fig. 6. Surface morphologies and cross-sections of AC, HR, and HR + CR samples of Zn–1Cu–0.1Ti after immersion in Hanks’ solution for 14 d and 30 d.

in the corrosion surfaces, indicating that the corrosion products 3.4. Friction and wear behavior of Zn–1Cu–0.1Ti
on the sample surfaces of Zn–1Cu–0.1Ti provide a protective effect
which prevents further corrosion. Fig. 8 shows the friction behaviors and Brinell hardness of the
Fig. 7 shows the EDS, XRD, FTIR, and XPS analysis results for AC, HR, and HR + CR samples of Zn–1Cu–0.1Ti and compared to
the corrosion products on the HR-processed Zn–1Cu–0.1Ti after im- AC pure Zn. The friction coefficient curves for Zn–1Cu–0.1Ti and
mersion in Hanks’ solution for 30 d. It can be seen that the corro- pure Zn during dry-sliding wear and corrosive wear testing in
sion products are mainly composed of C, O, P, Zn, Ca, and small Hanks’ solution are shown in Fig. 8(a). It can be seen that the fric-
amounts of Mg and Cu, as analyzed by EDS (Fig. 7(a)). The XRD tion coefficient of all the samples is relatively stable at the ini-
results reveal that the corrosion products consist of a Zn matrix, tial stage of wear under dry sliding, and then increases rapidly
ZnO, and ZnCO3 (Fig. 7(b)). In the FTIR spectrum of the corro- with sharp fluctuations with increasing wear time. The friction co-
sion products (Fig. 7(c)), vibration absorption peaks are observed at efficient of the AC pure Zn fluctuates between 0.695 and 1.541.
waves of approximately 2956, 1539, 1421, 1064, and 864 cm−1 , re- The friction coefficient of the AC Zn–1Cu–0.1Ti fluctuates between
spectively. According to the results reported in [48–51], a vibration 0.581 to 2.405; that of the HR and HR+CR Zn–1Cu–0.1Ti fluctuates
absorption peak at a wave in the range 1050–1180 cm−1 belongs between 0.587∼2.237 and 0.527∼2.411, respectively. Under friction
to PO34− ; in the ranges of 830–890 cm−1 , 1390–1500 cm−1 , and and wear conditions with solution lubrication, the friction coeffi-
1540–1620 cm−1 belongs to CO23− ; and in the range of 2920 cm−1 cients for all the samples show a similar trend, fluctuating signifi-
belongs to HPO24− . It can be concluded that PO34− , CO23− , and cantly in the initial stage of wear and then tending to become sta-
HPO24− co-exist in the corrosion products of the HR sample after ble after 5 min.
immersion in Hanks’ solution for 30 d. The XPS spectrum of the The Brinell hardness of the AC sample of pure Zn and AC, HR,
corrosion products on the HR-processed Zn–1Cu–0.1Ti after immer- and HR+CR samples of Zn–1Cu–0.1Ti is shown in Fig. 8(b). The
sion in Hanks’ solution for 30 d is shown in Fig. 7(d). Four distinct Brinell hardness of the studied samples was 28. 7 ± 0.1 HB for
peaks can be observed in the XPS spectrum, corresponding to the AC pure Zn, 71. 3 ± 0.3 HB for AC Zn–1Cu–0.1Ti, 70. 4 ± 0.2 HB
signal absorption peaks of the main elements including Zn, Ca, O, for HR Zn–1Cu–0.1Ti and 56. 2 ± 0.2 HB for HR+CR Zn–1Cu–0.1Ti,
and C, respectively. An XPS high-resolution narrow scan was per- respectively. It can be seen that the Brinell hardness of the stud-
formed on the main elements in the full spectrum and the results ied samples showed the same trend relative to the results of the
are shown in Fig. 7(e). The high-resolution narrow scanning spec- Vickers hardness shown in Fig. 4(b).
tra show values of 1044.8 and 1021.8 eV for Zn 2p1; 531.8 eV for O Based on the Brinell hardness results of the studied samples,
1 s; 347.4 eV for Ca 2p; and 285.0 eV for C 1 s. Furthermore, there it can be concluded that the wear loss of AC pure Zn and HR+CR
are C KLL broad peaks, Na 1 s, and Na KLL peaks in the full spec- Zn–1Cu–0.1Ti is higher than that of AC and HR Zn–1Cu–0.1Ti, as
trum (Fig. 7(d)). The Na peak might be derived from the deposition shown in Fig. 8(c).
of solute in Hanks’ solution. Based on the XPS data, it can be de- The friction coefficient, wear loss, and surface roughness of the
duced that the main components of the corrosion products on the AC sample of pure Zn and the AC, HR, and HR+CR samples of
HR sample are ZnO and ZnCO3 . Zn–1Cu–0.1Ti after dry-sliding wear and corrosive wear testing in
Combining the EDS, XRD, FTIR, and XPS results, it can be in- Hanks’ solution are shown in Fig. 8(c). The AC pure Zn shows a
ferred that the corrosion products of the HR-processed Zn–1Cu– friction coefficient of 0. 741 ± 0.037, a wear loss of 2. 41 ± 0.05 mg,
0.1Ti sample are mainly ZnO, ZnCO3 , and Zn3 (PO4 )2 ; and also pos- and a surface roughness of 1. 02 ± 0.05 μm; while the AC Zn–1Cu–
sibly ZnHPO4 because Hanks’ solution contains anions of HPO24− . 0.1Ti shows a friction coefficient of 1. 039 ± 0.024, a wear loss
Drelich et al. [52] reported that the corrosion products of pure Zn of 1. 87 ± 0.09 mg, and a surface roughness of 1. 19 ± 0.08 μm.
wires implanted in vivo over long-term surveillance were mainly After thermomechanical processes, the friction coefficient, wear
ZnO, ZnCO3 , and Zn3 (PO4 )2 , and their experimental results indi- loss, and surface roughness of the HR and HR+CR samples of Zn–
cated that pure Zn could be safely degraded in vivo within a time 1Cu–0.1Ti are 0. 796 ± 0.031, 1. 49 ± 0.03 mg, 0. 96 ± 0.08 μm,
frame of approximately 1∼2 years. and 0. 731 ± 0.052, 2. 02 ± 0.11 mg, 0. 94 ± 0.06 μm, respectively.
J. Lin, X. Tong and Z. Shi et al. / Acta Biomaterialia 106 (2020) 410–427 419

Fig. 7. EDS, XRD, FTIR, and XPS analysis results for corrosion products of HR samples of Zn–1Cu–0.1Ti after immersion in Hanks’ solution for 30 d: (a) EDS elemental
composition; (b) XRD spectra; (c) FTIR spectra; (d) XPS full spectrum; and (e) XPS high-resolution narrow scan results for Zn 2p, O 1 s, Ca 2p, and C 1 s.
420 J. Lin, X. Tong and Z. Shi et al. / Acta Biomaterialia 106 (2020) 410–427

Fig. 8. Friction behaviors and Brinell hardness of AC, HR, and HR + CR samples of Zn–1Cu–0.1Ti and AC pure Zn: (a) friction coefficient curves in dry-sliding wear and
corrosive wear testing in Hanks’ solution; (b) Brinell hardness; and (c) friction coefficient, wear loss and surface roughness after wear testing.

Under the dry friction and wear conditions, the friction coefficient, ter wear resistance in the Hanks’ solution environment than in the
wear loss, and surface roughness of the samples are much higher dry-sliding condition, and the HR sample of Zn–1Cu–0.1Ti shows
than those in the liquid lubrication environment. In addition, all much better wear resistance than the AC and HR+CR samples in
the studied alloy samples show a long, stable plateau in the liq- the Hanks’ solution environment.
uid lubrication environment. It can be concluded that the AC, HR, Fig. 9 shows the SEM images of the worn surface morphologies
and HR+CR samples of Zn–1Cu–0.1Ti and AC pure Zn exhibit bet- of the AC, HR, and HR+CR samples of Zn–1Cu–0.1Ti and AC pure

Fig. 9. Surface morphologies of AC, HR, and HR + CR samples of Zn–1Cu–0.1Ti and AC pure Zn after dry-sliding wear and corrosive wear tests.
J. Lin, X. Tong and Z. Shi et al. / Acta Biomaterialia 106 (2020) 410–427 421

0.1Ti alloy possesses good blood compatibility and meets the re-
quirements for the hemolysis rate of implant materials.

3.6. Cytotoxicity of Zn–1Cu–0.1Ti

The chemical composition, wettability, topography, and rough-


ness of biomaterial surfaces can affect cell compatibility [53].
Fig. 11 shows the contact angle and cell survival rate of the AC
pure Zn and Zn–1Cu–0.1Ti surfaces. It can be seen that the contact
angles of the AC pure Zn and Zn–1Cu–0.1Ti are 62. 4 ± 0.9° and
63. 7 ± 2.6°, respectively (Fig. 11(a) and (b)). This result indicates
hydrophilicity and there is no noticeable difference between the
contact angles of both surfaces. Solař et al. [54] reported that hy-
drophilic surfaces have better cell adhesion than hydrophobic sur-
faces, and Li et al. [55] reported that the cells tend to adhere onto
surfaces with a water contact angle of approximately 70°. There-
fore, the surface contact angles of both pure Zn and Zn–1Cu–0.1Ti
should be beneficial to cell adhesion and attachment.
Fig. 10. Hemolysis percentages of AC pure Zn and Zn–1Cu–0.1Ti samples. The survival rates of MC3T3-E1 pre-osteoblast cells after cul-
turing on AC pure Zn and Zn–1Cu–0.1Ti for 24 and 48 h are
shown in Fig. 11(c). As can be seen, the survival rate of MC3T3-
Zn after dry-sliding wear and corrosive wear testing in Hanks’ so- E1 cells after culturing for 24 h is 73. 6 ± 1.0% for AC pure Zn and
lution. The worn surface of the samples exhibit many plowed fur- 44. 3 ± 3.3% for AC Zn–1Cu–0.1Ti. When the culturing time reaches
rows under dry-sliding friction which were wide and deep, and 48 h, the survival rate decreases to 55. 5 ± 8.4% for AC pure Zn and
parallel to the sliding direction. The edges of the plowed furrows 13. 9 ± 0.6% for AC Zn–1Cu–0.1Ti. The survival rates of the cells af-
show visible plastic deformation and some pits are formed in lo- ter culturing for 24 and 48 h are lower than 75%, showing a cyto-
calized areas due to the metal falling off. After friction and wear in toxicity grade of 2–4 as defined by ISO 10993-5 [39].
Hanks’ solution, all the samples show smooth worn surfaces, and Fig. 12 shows the ion concentrations of the extracts of AC pure
there are shallow and narrow plowed furrows. The wear mech- Zn and Zn–1Cu–0.1Ti and the cell viability of MC3T3-E1 and MG-
anism consists mainly of abrasive wear and almost no adhesive 63 cells after culturing with different concentrations of extracts
wear. In addition, there is a mixture of corrosion and wear debris for 1, 3, and 5 d. It can be seen that the Zn ion concentration
on the outer edges of the wear scars. The wear mechanism dur- in the extract of AC pure Zn is 14. 16 ± 0.26 μg/ml and the Zn,
ing dry-sliding wear consists of adhesive wear and plastic deforma- Cu, and Ti ion concentrations in the extract of Zn–1Cu–0.1Ti are
tion, and the wear mechanism under solution lubrication consists 33. 97 ± 2.11 μg/ml, 14. 2 ± 0.9 ng/ml, and 0. 4 ± 0.2 ng/ml, re-
of plowing and slightly abrasive wear. In addition, the width of the spectively (Fig. 12(a)). The undiluted extracts (i.e., 100% concentra-
grinding surface after dry-sliding wear and wet lubrication wear tion) of AC pure Zn and Zn–1Cu–0.1Ti show low cell viability for
in descending order is AC pure Zn > HR+CR Zn–1Cu–0.1Ti > AC both MC3T3-E1 and MG-63 cells at all culturing time points of 1,
Zn–1Cu–0.1Ti > HR Zn–1Cu–0.1Ti. 3, and 5 d (Fig. 12(b)). The cell viability is lower than 55% and
the extracts are substantially cytotoxic due to the high metal ion
3.5. Hemolysis percentages of Zn–1Cu–0.1Ti concentrations. When the concentration of the extracts of both AC
pure Zn and Zn–1Cu–0.1Ti is diluted to 25%, the corresponding cell
Fig. 10 shows the hemolysis percentages of the AC samples of viability significantly increases after culturing for all time points
Zn–1Cu–0.1Ti compared to the AC pure Zn. It can be seen that the (Fig. 12(c)). The cell viability of MC3T3-E1 cells is 64. 48 ± 1.02%
hemolysis percentages of red blood cells are 2. 44 ± 0.08% and for pure Zn and 82. 74 ± 0.34% for Zn–1Cu–0.1Ti after culturing
3. 62 ± 0.14% for the extracts of pure Zn and Zn–1Cu–0.1Ti al- for 1 d The cell viability of MG-63 cells is 87. 30 ± 1.98% for pure
loy, respectively. According to the requirements for the hemolysis Zn and 65. 45 ± 0.42% for Zn–1Cu–0.1Ti after culturing for 1 d Af-
rate of ISO 10993-4: 2002 [38], the hemolysis percentages of clini- ter further diluting the extracts of AC pure Zn and Zn–1Cu–0.1Ti to
cally applicable implant materials should less than 5%. The hemol- 12.5% concentration, the cell viability of both MC3T3-E1 and MG-
ysis percentages of both these materials are far below 5% and no 63 cells exceeds 90% after culturing for 1 d, indicating good in vitro
hemolysis occurs. The hemolysis results indicate that the Zn–1Cu– cytocompatibility and non-toxicity. The cell viability with different

Fig. 11. Contact angles and cell survival rates of AC Zn and Zn–1Cu–0.1Ti surfaces: (a) contact angle of pure Zn; (b) contact angle of Zn–1Cu–0.1Ti; and (c) survival rate of
MC3T3-E1 cells after culturing for 24 and 48 h.
422 J. Lin, X. Tong and Z. Shi et al. / Acta Biomaterialia 106 (2020) 410–427

Fig. 12. (a) Ion concentrations of extracts of AC pure Zn and Zn–1Cu–0.1Ti; cell viability of MC3T3-E1 and MG-63 cells after culturing with extracts of AC pure Zn and
Zn–1Cu–0.1Ti for 1, 3, and 5 d at concentrations of: (b) 100%; (c) 25%; and (d) 12.5%.

extract concentrations shows a similar trend after culturing for 3 by adding the alloying elements of Cu and Ti occurred through var-
and 5 d, and the cell viability of MC3T3-E1 and MG-63 cells in- ious mechanisms including Hall-Petch strengthening, solid-solution
creases with the dilution of the extracts. Furthermore, the cell via- strengthening, and precipitation hardening. Firstly, due to the ad-
bility of MC3T3-E1 and MG-63 cells cultured with the extracts for dition of the Cu and Ti alloying elements, the η-Zn matrix phase in
1, 3, and 5 d shows no significant difference between AC pure Zn the AC Zn–1Cu–0.1Ti showed an average grain size of ∼54 μm, sig-
and Zn–1Cu–0.1Ti. nificantly less than that of the AC pure Zn (∼600 μm). The grain
size of the AC Zn–1Cu–0.1Ti is over 11 times finer than that of
3.7. Antibacterial property of Zn–1Cu–0.1ti the AC pure Zn. It is well known that metals of finer grain size
exhibit greater yield strength based on the Hall-Petch strengthen-
Fig. 13 shows the antibacterial effect of AC pure Zn and Zn– ing mechanism (that is, σ y = σ o + k d−1/2 , where σ y is the yield
1Cu–0.1Ti after co-culturing with S. aureus for 24 h at 37 °C. A bac- stress, σ o is a materials constant for starting stress for dislocation
teriostatic ring appears around the samples of pure Zn and Zn– movement (or resistance of the lattice to dislocation motion), k is
1Cu–0.1Ti, and the value of IZD (H) is 6. 19 ± 0.42 mm for pure Zn the strengthening coefficient (a constant unique to each material),
and 6. 99 ± 0.33 mm for Zn–1Cu–0.1Ti. The Zn–1Cu–0.1Ti exhibits and d is the grain diameter) [56–57]. Secondly, the addition of the
a good antibacterial effect and its IZD is higher than that of pure Cu element resulted in solid-solution strengthening of the Zn alloy
Zn, and the differences between the two groups were statistically matrix. Some of the Cu atoms and the η-Zn matrix phase formed
significant (P < 0.05). a solid-solution phase because the difference of their atomic radii
is less than 15%, causing lattice strains on surrounding host atoms
4. Discussion which provide resistance for dislocation movement, leading to en-
hanced strength and hardness relative to those of the pure Zn [58].
4.1. Microstructures and mechanical properties of Zn–1Cu–0.1Ti Third, the remaining Cu and Ti formed fine and reticular second
phases of ε -CuZn5 and TiZn16 distributed along the grain bound-
The AC Zn–1Cu–0.1Ti exhibits σ uts = 92.4 MPa, σ ys = 86.1 MPa, aries or inside the grains. These second phases resulted in precip-
ε = 1.4%, and hardness of 72.6 HV, respectively, significantly higher itation hardening since the particles resist the dislocation motion
than those of AC pure Zn (σ uts = 33.6 MPa, σ ys = 29.3 MPa, and developing the bow around the second particles, which led to
ε = 1.2%, and hardness = 37.9 HV [18]), indicating that the ad- increased hardness of the Zn alloy.
dition of Cu and Ti can effectively improve the tensile properties As for the HR sample of Zn–1Cu–0.1Ti, deformation heat ac-
and hardness of the Zn alloy. Strengthening of Zn–1Cu–0.1Ti alloy cumulated during the cyclical heating and deformation of the
J. Lin, X. Tong and Z. Shi et al. / Acta Biomaterialia 106 (2020) 410–427 423

Fig. 13. Antibacterial effect of AC pure Zn and Zn–1Cu–0.1Ti after co-culturing with S. aureus for 24 h: (a) image of pure Zn; (b) image of Zn–1Cu–0.1Ti; and (c) inhibition
zone diameter (IZD) of pure Zn and Zn–1Cu–0.1Ti.

hot-rolling process, which caused localized dynamic recoveries and to the difference in their corrosion potentials. The precipitated
recrystallization due to the low recrystallization temperature of the phase serves as the cathode electrode and a potential difference
alloy. This led to a significant dynamic softening of the alloy and is formed between the precipitated particles and the matrix phase,
thus the hardness of the HR sample decreased slightly and the causing galvanic corrosion which accelerates the intergranular cor-
elongation increased notably as compared to the AC sample, while rosion of the Zn alloy. Secondly, they exert a barrier effect. When
the fracture mode changed from a brittle cleavage fracture to a the second phase of a Zn alloy is distributed in the matrix and
ductile dimple fracture. forms a continuous network, it can act as a corrosion barrier for
In the case of the HR+CR sample of Zn–1Cu–0.1Ti, more ε - the Zn matrix, preventing the corrosion from proceeding fully and
CuZn5 particles precipitated during the hot-rolling and cold-rolling thereby slowing the corrosion of the Zn alloy.
processes in the alloy, which resulted in a reduced concentra- The microstructure of the AC Zn–1Cu–0.1Ti is composed of an
tion of Cu in the η-Zn matrix phase, weakening the solid-solution η-Zn phase and the second phases of TiZn16 and ε-CuZn5 . These
strengthening of the Cu. This caused a significant decrease in hard- two second phases are mainly distributed along the grain bound-
ness. Moreover, the hard and brittle ε -CuZn5 particles that precip- aries and form a network in the matrix, and only some of them are
itated were crushed into fine particles during the hot rolling and distributed inside the grains, so the network of the second phases
cold rolling. These fine ε -CuZn5 particles homogeneously dispersed can act as a barrier and hinder corrosion. As for the HR sample of
in the η-Zn matrix, leading to significantly increased strength and Zn–1Cu–0.1Ti, the TiZn16 intermetallic phase of the network in the
ductility. Therefore, the HR+CR sample of Zn–1Cu–0.1Ti exhibits AC alloy was crushed and formed a fiber-like or dispersed gran-
the best overall mechanical properties among the three samples. ular structure along the rolling direction, resulting in a weaken-
ing of the barrier effect. In such cases galvanic corrosion plays a
4.2. Corrosion behavior of Zn–1Cu–0.1Ti major role, resulting in an accelerated corrosion rate of the alloy
with a decreased corrosion resistance compared to the AC alloy. Af-
The standard electrode potential is −0.763 V for Zn, 0.342 V for ter HR processes, the coarse skeletal ε -CuZn5 phase of the HR+CR
Cu, and −1.630 V for Ti [56,59]. Therefore, alloying with Cu can be Zn–1Cu–0.1Ti is refined and dispersed in the matrix, resulting in
expected to increase the corrosion potential of a Zn alloy due to a significant reduction of the active cathode region and weakened
its more positive standard electrode potential than that of the Zn. galvanic corrosion, therefore exhibiting better corrosion resistance.
However, the corrosion potential of the η-Zn phase matrix is lim- According to Ren et al. [61], the degradation process of pure
ited due to the low content of Cu. Both Cu and Ti mainly form Zn and Zn–Mg–Zr alloys in SBF can be expressed by the reaction
the second phases of ε -CuZn5 , TiZn16 , and probably Cu2 TiZn22 and between Zn and water:
other precipitation phases with Zn. In general, second phases have
Zn + H2 O → ZnO + 2H+ + 2e− (4)
a dual effect on the corrosion behavior of the relevant material
[60]. Firstly, they cause galvanic corrosion. The second phase and When the immersion time is short (14 d), the corrosion prod-
the matrix phase of a Zn alloy will form a micro-battery cell due uct adheres to the surface of the sample in a point-like manner
424 J. Lin, X. Tong and Z. Shi et al. / Acta Biomaterialia 106 (2020) 410–427

without forming a dense corrosion layer. It forms galvanic corro- of wear debris that is not easily discharged increases with testing
sion with the η-Zn phase matrix in which the η-Zn phase acts time and the obstruction effect on the friction pair is enhanced, re-
as the anode and the corrosion product as the cathode. This leads sulting in severe abrasive wear on the surface of the samples. The
to high degradation of all the samples of AC, HR, and HR+CR Zn– friction coefficient of the samples in dry-sliding wear increases and
1Cu–0.1Ti. With a prolonged immersion time (30 d), the corrosion fluctuates substantially, indicating a type of three-body wear.
product layer gradually thickens and becomes denser, which can In addition, there is a clear running-off period in the case of
prevent the Cl− ions in the environment from entering the ma- dry-sliding wear, while the running-in period is short or even ab-
trix and so prevent further corrosion to a certain extent, thereby sent in the case of wet friction and wear with Hanks’ solution.
slowing the corrosion rate and protecting the substrate of the al- This is mainly because dry friction wear is a type of three-body
loy. Therefore, the studied alloy samples show a lower degradation wear and its main wear mechanism is a fatigue layer undergoing
rate after immersion in Hanks’ solution for 30 d. fatigue fracture due to the multiple plastic deformation, while the
wet friction and wear is a type of two-body abrasive wear, which
4.3. Friction and wear behavior of Zn–1Cu–0.1Ti does not require a long running-in period, and so the stable pe-
riod becomes longer. Therefore, for the abovementioned reasons,
As can be seen from Fig. 8(c), the AC sample of pure Zn and the the friction coefficient, wear loss, and surface roughness of all the
HR+CR sample of Zn–1Cu–0.1Ti shows a lower friction coefficient samples under the dry friction condition are significantly higher
and higher wear loss and surface roughness than the AC and HR than those under the wet friction and wear environment of Hanks’
Zn–1Cu–0.1Ti. This is because the hardness of both samples was solution.
considerably lower than that of the AC and HR Zn–1Cu–0.1Ti.
According to the Holm-Archard equation [62], the wear loss Q 4.4. Hemolytic properties, cytocompatibility, and antimicrobial ability
can be calculated using the formula given by: of Zn–1Cu–0.1Ti
Q = kw/H (5)
New biodegradable implant materials for bone fixation not only
where k is the wear factor, w is the applied load, and H is the need to exhibit suitable biodegradability and mechanical prop-
hardness of the worn surface. When the applied load is a constant erties, but also require good biocompatibility and even biofunc-
200 gf (∼2 N), the wear rate is inversely proportional to the hard- tionalities such as antibacterial ability. The recommended daily
ness. According to the macro-hardness data of Brinell hardness for intakes of Zn, Cu, and Ti ions for adults are about 12∼16 mg,
the samples shown in Fig. 8(b), it can be concluded that the wear 0.9∼1.2 mg, and 0.3∼2 mg, respectively and excessive intake of
loss of the AC pure Zn and HR+CR Zn–1Cu–0.1Ti is higher than these metal ions causes systemic toxicity [49,64–66]. Most of the
that of the AC and HR Zn–1Cu–0.1Ti, indicating the same trend as Zn, Cu, and Ti ions absorbed into the systemic circulation are ex-
shown in Fig. 8(c). creted through feces and urine. As shown in Fig. 5(f), the degra-
The AC sample of pure Zn and the AC, HR, and HR+CR sam- dation rate of HR Zn–1Cu–0.1Ti after immersion in Hanks’ solu-
ples of Zn–1Cu–0.1Ti exhibit superior wear resistance in the liquid tion for 30 d is 0.032 mm/y. The release efficiency of each kind
environment of Hanks’ solution than in the dry-sliding wear envi- of ion after implantation can be estimated according to the cor-
ronment. This is because the Hanks’ solution acts in three ways to responding degradation rate, and the corrosion rate can be cal-
affect the friction and wear of the samples during corrosive wear. culated to be 0.016 mg/cm2 /day. Assuming that the dimensions of
Firstly, it provides a lubrication effect on the samples during the a screw implant are 2.5 mm × 18 mm with a surface area of ap-
corrosive-wear process and the solution acts as a lubricant. The proximately 2.0 cm2 [67], it can be estimated that the ion release
hard abrasive particles falling off the surface of the samples are rates of Zn, Cu, and Ti are 32.4 μg/d, 0.33 μg/d, and 0.03 μg/d, re-
washed away by the solution, so these hard particles can no longer spectively, which is far below the recommended daily intake levels
abrade the sample. The wear mechanism under the lubrication of of these elements. In addition, Bian et al. [49] reported that the
Hanks’ solution is therefore slight abrasive wear, which is a type local ion concentrations of Mg and Ge surrounding the implant in
of two-body wear and the change of friction coefficient during the vivo were diluted and much lower than those of the in vitro testing
wear process is relatively smooth and stable. Secondly, the Hanks’ experimental results due to the circulation system. In addition, be-
solution provides cooling to the process. At a constant temperature cause the corrosion products of the Zn–1Cu–0.1Ti alloy in Hanks’
of 37 °C, Hanks’ solution has a good cooling effect on the friction solution are mainly composed of ZnO, ZnCO3 , and ZnPO3 , these
surface, which can take away the heat generated by the friction products are similar to the degradation products of the Mg-based
process over time and reduce the tendency of the alloy to undergo alloy implants reported by Witte et al. [68], which were nontoxic
adhesive wear. Thirdly, although the Hanks’ solution has a corro- and non-irritating to the human body. It can be concluded that the
sive effect on the Zn alloy, Zn alloys possess much better corrosion degradation products of the Zn–1Cu–0.1Ti alloy can be excreted by
resistance than Mg alloys and so the corrosion effect on the Zn al- the physiological system and should not cause systemic toxicity.
loy is insignificant in a short wear period. Therefore, the medium When the AC pure Zn and Zn–1Cu–0.1Ti samples are directly
of the Hanks’ solution does not aggravate the wear on the Zn al- cultured with MC3T3-E1 and MG-63 cells, both samples show low
loy, unlike the aggravated wear on Mg alloys in SBF reported by cell viability, indicating that the surfaces of both materials produce
Liu et al. [63]. certain toxic effects. This is because the concentration of the metal
In the dry-sliding friction process, since the hardness of the ions released is higher than the threshold of the ions required for
friction pair is relatively high, the hardness difference between the cell proliferation, resulting in a decrease in the proliferation capac-
test samples and the friction pair is relatively significant. So as the ity of the MC3T3-E1 cells. In addition, the Zn–1Cu–0.1Ti alloy ex-
wear time is considerably lengthened, the surface temperature of hibited lower cell activity than pure Zn, mainly because its degra-
the alloy continuously increases due to the accumulation of friction dation rate (314.6 μm/y) is faster than that of pure Zn (156.7 μm/y
heat, resulting in partially softened surfaces of the metal samples for AC pure Zn obtained from potentiodynamic polarization tests
which leads to the adhesive wear and plastic deformation. In ad- in Hanks’ solution [18]). The excess metal ions are released from
dition, the partially peeled protective layers exists in the form of the surface of the sample during cell culturing, which causes pro-
solid debris between the friction pair and samples, and the abrad- found cell death. With culture time increasing, the concentration
ing effect of the debris on the samples causes plowed furrows of metal ions in the culture solution gradually increases, resulting
whose depth and width increase with testing time. The amount in a continuous decrease in the survival rate of MC3T3-E1 cells. In
J. Lin, X. Tong and Z. Shi et al. / Acta Biomaterialia 106 (2020) 410–427 425

response to this, many researchers [20,21] proposed using diluted 2. The HR+CR sample of the Zn–1Cu–0.1Ti alloy exhibited the
alloy extracts to evaluate the in vitro cytotoxicity of metallic bio- highest tensile properties among the three conditions of AC,
materials. HR, and HR+CR, with a yield strength of 204.2 MPa, an ulti-
To assess the effect of extract concentration on in vitro cyto- mate strength of 249.9 MPa, and an elongation of 75.2%; these
compatibility tests, Wang et al. [69] recommended that cytotoxic- are 2.7 times the yield strength, 2.4 times the ultimate ten-
ity be assessed using a diluted extract method, which use a 6∼10 sile strength, and 53.7 times the elongation of the AC alloy. The
times dilution of extracts for in vitro cytotoxicity testing. Dambatta fracture mode of the Zn–1Cu–0.1Ti alloy was transformed from
et al. [70] found that osteoblast cells were highly active when cul- brittle fractures with mixed fracture characteristics of river pat-
tured in extracts of a Zn–3Mg alloy with a Zn ion concentration terns, cleavage steps, and a few small dimples in localized ar-
below 0.5 mg/ml (0.1 and 0.5 mg/l), but more toxic when cultured eas in the AC alloy to ductile fractures with homogeneously
in extracts with ion concentrations of 1.0 mg/ml and 2.0 mg/ml. distributed, round, equiaxed dimples in the HR alloy and the
Furthermore, when osteoblast cells were cultured in alloy extracts HR+CR alloy. The dimples on the fracture surface of the HR+CR
with Zn ion concentrations of 0.1 and 0.5 mg/ml for 1, 3, and 7 d, alloy were deeper, larger, and denser compared to those of the
the relative activity of cells increased to the maximum and then HR alloy. The HR+CR alloy showed a hardness of 56.4 HV, lower
decreased with the elongation of culture time. Ma et al. [71] also than those of the HR alloy (70.6 HV) and the AC alloy (72.6 HV).
found that extracts with low concentrations of Zn ions promoted 3. Potentiodynamic polarization tests on Zn–1Cu–0.1Ti showed
cell proliferation and growth, but extracts with high concentrations that the corrosion potential, corrosion current density, and cor-
of Zn ions inhibited cell activity. For the difference of cell viabil- rosion rate were −1.025 V, 21.5 μA/cm2 , and 315 μm/y for the
ity between the MC3T3 and MG63 cells, it may be because the AC alloy; −1.123 V, 111.2 μA/cm2 , and 1628 μm/y for the HR al-
cell proliferation and tolerance of the both cell lines for Zn ions loy; and −1.100 V, 67.7 μA/cm2 , and 991 μm/y for the HR+CR
with different level were different. This kind of phenomenon has alloy. The degradation rate of Zn–1Cu–0.1Ti after immersion
also been reported by other researchers, for example, Shen et al. in Hanks’ solution for 30 d was 0.029 mm/y, 0.032 mm/y, and
[72] reported that the HOS cells showed a significant reduction in 0. 034 ± 0.018 mm/y for the AC, HR and HR+CR alloys, re-
cell viability and induced cytotoxicity at a higher extract concen- spectively. The corrosion products consisted of Zn(CO)3 , ZnO,
tration; while the MG63 cells had a slight negative impact on cell Zn3 (PO4 )2 , and also probably ZnHPO4 .
viability at the same level of extract. 4. The AC, HR, and HR+CR samples of Zn–1Cu–0.1Ti and the AC
In this study, the relative cell viability of both MC3T3-E1 and pure Zn exhibited better wear performances in the liquid lubri-
MG-63 cells increases as the concentration of the extract decreases cation environment of Hanks’ solution than in the dry-sliding
(Fig. 12). Moreover, the results indicate that a low concentration wear condition, and the HR sample showed significantly better
extract (≤ 25%) of Zn–1Cu–0.1Ti shows no apparent toxic effect on wear resistance among the samples. The wear mechanisms of
MC3T3-E1 cells. the dry-sliding wear and corrosion wear were three-body abra-
For the effect of metal ions on antibacterial activity, Sukhodub sive wear and two-body abrasive wear, respectively.
et al. [73] systematically studied the antibacterial ability of metal 5. The extract of AC Zn–1Cu–0.1Ti alloy at a concentration ≤ 25%
ions and reported that the metal ion sterilization rate from high showed no significant cytotoxicity to MC3T3-E1 and MG-63
to low was: Ag+ , Cu2+ , Zn2+ , Mg2+ . Among these metal ions, Zn cells. The AC Zn–1Cu–0.1Ti alloy exhibited good blood compat-
ions have a good antibacterial ability when they reach a certain ibility and antibacterial property.
concentration and can kill various bacteria and fungi. Wahab et al.
[74] reported that a concentration of ≥ 11 mM of Zn acetate had Overall, based on the testing results for mechanical proper-
a good killing effect on S. aureus. Cu and its alloys also had good ties, corrosion and corrosive-wear property, degradation behavior,
bacteria-killing effects on pathogenic bacteria such as Escherichia biocompatibility, and antibacterial ability, the Zn–1Cu–0.1Ti alloy
coli, Mycobacterium tuberculosis, and methicillin-resistant S. au- is a promising candidate for bone fracture fixation applications
reus [75–77]. Tang et al. [21] reported that the antibacterial proper- such as bone screws and plates of biodegradable implants due to
ties of Zn–xCu (x = 1, 2, 3, 4%) alloys increased with an increase in its markedly enhanced mechanical properties, wear resistance, an-
Cu content. The antibacterial testing in this study shows that the tibacterial ability, and suitable degradation rate.
Zn–1Cu–0.1Ti alloy has a good antibacterial ability which is better
than that of pure Zn, as indicated by a higher IZD (Fig. 13). Declaration of Competing Interest

5. Conclusions The authors declare that they have no known competing finan-
cial interests or personal relationships that could have appeared to
This Zn–1Cu–0.1Ti alloy has been developed as a biodegradable influence the work reported in this paper.
Zn-based alloy for bone implant applications. It achieves enhanced
mechanical properties (both strength and ductility) and antibacte- Acknowledgments
rial ability through alloying Zn with Ti and Cu. Thermomechanical
processes of HR and HR+CR were performed on the AC Zn–1Cu– This work was supported by the Wenzhou Science and Technol-
0.1Ti alloy to further boost the mechanical properties. The main ogy Bureau through the project G20180027 and 2018ZG008. CW
conclusions are as follows: and YL also acknowledge the financial support for this research
by the Australian Research Council (ARC) through the Discovery
1. The microstructure of the AC Zn–1Cu–0.1Ti alloy was composed
Project DP170102557 and Future Fellowship FT160100252. JL ac-
of a η-Zn matrix phase, an intermetallic compound TiZn16
knowledges the support for this research by Hunan Provincial Nat-
phase, and an ε -CuZn5 phase. Most of the intermetallic com-
ural Science Foundation of China through the Grant 2016JC2005.
pound TiZn16 phase was distributed along the grain boundaries
of the alloy. The alloy after HR exhibited elongated η-Zn, TiZn16 ,
and Cu2 TiZn22 phases along the rolling direction. The alloy after References
HR+CR exhibited visible processing flow lines and finely dis-
[1] D.M. Zhang, D.X. Cui, R.S. Xu, Y.C. Zhou, L.W. Zheng, P. Liu, X.D. Zhou, Pheno-
persed ε -CuZn5 particles in the η-Zn matrix along the rolling typic research on senile osteoporosis caused by SIRT6 deficiency, Int. J. Oral
direction. Sci. 8 (2016) 84–92.
426 J. Lin, X. Tong and Z. Shi et al. / Acta Biomaterialia 106 (2020) 410–427

[2] M.G. Oefelein, V. Ricchuiti, W. Conrad, A. Seftel, D. Bodner, H. Goldman, (Ti–12Mo–6Zr–2Fe) in simulated body fluid, Acta Biomater. 42 (2016)
M. Resnick, Skeletal fracture associated with androgen suppression induced 429–439.
osteoporosis: the clinical incidence and risk factors for patients with prostate [32] P.E. Sinnett-Jones, J.A. Wharton, R.J.K. Wood, Micro-abrasion-corrosion of a
cancer, J. Urol 166 (5) (2001) 1724–1728. CoCrMo alloy in simulated artificial hip joint environments, Wear 259 (2005)
[3] M.P. Staiger, A.M. Pietak, J. Huadmai, G. Dias, Magnesium and its alloys as or- 898–909.
thopedic biomaterials: a review, Biomaterials 27 (9) (2006) 1728–1734. [33] D.B. Liu, B. Wu, X. Wang, M.F. Chen, Corrosion and wear behavior of
[4] J.E. Gray-Munro, C. Seguin, M. Strong, Influence of surface modification on the an Mg–2Zn–0.2Mn alloy in simulated body fluid, Rare Met. 34 (8) (2015)
in vitro corrosion rate of magnesium alloy AZ31, J. Biomed. Mater. Res., Part A 553–559.
91 (1) (2008) 221–230. [34] ASTM E8/E8M-08, Standard test methods for tension testing of metallic mate-
[5] Y. Ding, C.E. Wen, P. Hodgson, Y.C. Li, Effects of alloying elements on the corro- rials, ASTM International, West Conshohocken, PA, 2008.
sion behavior and biocompatibility of biodegradable magnesium alloys: a re- [35] A. Biesiekierski, D.H. Ping, Y. Yamabe-Mitarai, C.E. Wen, Impact of ruthe-
view, J. Mater. Chem. B 2 (14) (2014) 1912–1933. nium on microstructure and corrosion behavior of β -type Ti–Nb–Ru alloys for
[6] M. Schinhammer, A.C. Hänzi, J.F. Löffler, P.J. Uggowitzer, Design strategy for biomedical applications, Mater. Des. 59 (2014) 303–309.
biodegradable Fe-based alloys for medical applications, Acta Biomater 6 (5) [36] ASTM G102-89, Standard practice for calculation of corrosion rates and related
(2010) 1705–1713. information from electrochemical measurements, ASTM International, West
[7] C.M. Bünger, N. Grabow, K. Sternberg, M. Goosmann, K.P. Schmitz, Conshohocken, PA, 2015.
H.J. Kreutzer, H. Ince, S. Kische, C.A. Nienaber, D.P. Martin, S.F. Williams, E. Klar, [37] ASTM G31-72, Standard Practice For Laboratory Immersion Corrosion Testing
W. Schareck, A biodegradable stent based on poly (L-lactide) and poly (4-hy- of Metals, ASTM International, Philadelphia West Conshohocken, PA, 2004.
droxybutyrate) for peripheral vascular application: preliminary experience in [38] ISO10993-4: 2002/ Amd 1:2006, Biological Evaluation of Medical Devices-Part
the pig, J. Endovasc. Ther. 14 (5) (2007) 725–733. 4: Selection of Tests For Interactions With Blood, International Organization
[8] J. Franco, A. Souto-Serantes, P.R. Rodriguez, F. Guitián, A.M. Ínsua, Ceramic pro- for Standardization, ANSI/AAMI, Arlington, VA, 2006.
cessing of beta-TCP for producing implantable shapes, Bol. Soc. Esp. Ceram. V. [39] ISO 10993-5: 2009, Biological Evaluation of Medical Devices-Part 5: Tests for in
45 (4) (2006) 265–270. Vitro Cytotoxicity, International Organization for Standardization, ANSI/AAMI,
[9] G. Ryan, A. Pandit, D.P. Apatsidis, Fabrication methods of porous metals for use Arlington, VA, 2009.
in orthopaedic applications, Biomaterials 27 (13) (2006) 2651–2670. [40] DIN EN ISO 20645-2004, Textile Fabrics-Determination of Antibacterial Activ-
[10] C. Hidaka, S. Maher, J. Packer, S. Gasinu, M.E. Cunningham, S. Rodeo, ity-Agar Diffusion Plate Test, International Organization for Standardization,
What’s new in orthopaedic research, J. Bone. Joint. Surg. Am. 91 (11) (2009) 2004.
2756–2770. [41] M. Saillard, G. Develey, C. Becle, J.M. Moreau, D. Paccard, The structure of
[11] H. Guo, R.H. Cao, Y.F. Zheng, J. Bai, F. Xue, C.L. Chu, Diameter-dependent in vitro TiZn16 , Acta Crystallogr 37 (1) (2001) 224–226.
performance of biodegradable pure Zn wires for suture application, J. Mater. [42] G. Boczkal, Controlling the morphology and distribution of an intermetallic
Sci. Technol. 35 (8) (2019) 1662–1670. Zn16 Ti phase in single crystals of Zn–Ti, Mod. Aspects Bulk Cryst, Thin. Film.
[12] M. Ezechieli, M. Ettinger, C. König, A. Weizbauer, P. Helmecke, R. Schavan, Prep. (2012) 141–162.
C. Becher, Biomechanical characteristics of bioabsorbable magnesium-based [43] Z. Zermout, M. Durand-Charre, G. Kapelski, B. Baudelet, Phase equilibria in the
(MgYREZr-alloy) interference screws with different threads, Knee Surg. Sport Zn-rich corner of the Zn–Cu–Ti system, Z. Metallkd 87 (1996) 274–279.
Tr. A 24 (12) (2016) 3976–3981. [44] T.B. Massalski, H. Okamoto, P.R. Subramanian, L. Kacprzak, Binary Alloy Phase
[13] C.G. Fraga, Relevance, essentiality and toxicity of trace elements in human Diagrams, ASM International, 1990.
health, Mol. Aspects Med. 26 (4–5) (2005) 235–244. [45] F. Rosalbino, G. Scavino, D. Macciò, A. Saccone, Influence of the alloying com-
[14] H. Tapiero, D.M. Townsend, K.D. Tew, Trace elements in human physiology ponent on the corrosion behaviour of Zn in neutral aerated sodium chloride
and pathology: Zn and metallothioneins, Biomed. Pharmacother 57 (2003) solution, Corros. Sci. 89 (12) (2014) 286–294.
399–411. [46] Y. Liu, H.Y. Li, Z.H. Li, EIS investigation and structural characterization of dif-
[15] K.D. Cashman, Milk minerals (including trace elements) and bone health, Int. ferent hot-dipped Zn-based coatings in 3.5% NaCl solution, Int. J. Electrochem.
Dairy J. 16 (2006) 1389–1398. Sci. 8 (6) (2013) 7753–7767.
[16] K. Kaur, R. Gupta, S.A. Saraf, S.K. Saraf, Zn: the metal of life, Compr. Rev. Food [47] Z.B. Wang, H.X. Hu, C.B. Liu, Y.G. Zheng, The effect of fluoride ions on the cor-
Sci. Food Saf. 13 (2014) 358–376. rosion behavior of pure titaniumin 0.05 M sulfuric acid, Electrochim. Acta 135
[17] J.Z. Ilich, J.E. Kerstetter, Nutrition in bone health revisited: a story beyond cal- (2014) 526–535.
cium, J. Am. Coll. Nutr. 19 (20 0 0) 715–737. [48] Y. Wu, S. Wen, K. Chen, J. Wang, G. Wang, K. Sun, Enhanced corrosion resis-
[18] X. Tong, D.C. Zhang, X.T. Zhang, Y.C. Su, Z.M. Shi, K. Wang, J.G. Lin, Y.C. Li, tance of waterborne polyurethane containing sulfonated graphene/zinc phos-
J.X. Lin, C.E. Wen, Microstructure, mechanical properties, biocompatibility, phate composites, Prog. Org. Coat. 132 (2019) 409–416.
and in vitro corrosion and degradation behavior of a new Zn–5Ge alloy for [49] D. Bian, W.R. Zhou, J.X. Deng, Y. Liu, W.T. Li, X. Chu, P. Xiu, H. Cai, Y.H. Kou,
biodegradable implant materials, Acta Biomater. 82 (2018) 197–204. B.G. Jiang, Y.F. Zheng, Development of magnesium-based biodegradable metals
[19] H.F. Li, X.H. Xie, Y.F. Zheng, Y. Cong, F.Y. Zhou, K.J. Qiu, X. Wang, S.H. Chen, with dietary trace element germanium as orthopaedic implant applications,
L. Huang, L. Tian, L. Qin, Development of biodegradable Zn–1X binary alloys Acta Biomater. 64 (2017) 421–436.
with nutrient alloying elements Mg, Ca and Sr, Sci. Rep. 5 (2015) 10719. [50] P.K. Bowen, J. Drelich, J. Goldman, Magnesium in the murine artery: probing
[20] J. Kubásek, D. Vojtěch, E. Jablonská, I. Pospíšilová, J. Lipov, T. Ruml, Structure, the products of corrosion, Acta Biomater. 10 (3) (2014) 1475–1483.
mechanical characteristics and in vitro degradation, cytotoxicity, genotoxicity [51] A.F. Anvari-Yazdi, K. Tahermanesh, S.M.M. Hadavi, T. Talaei-Khozani,
and mutagenicity of novel biodegradable Zn–Mg alloys, Mater. Sci. Eng., C 58 M. Razmkhah, S.M. Abed, M.S. Mohtasebi, Cytotoxicity assessment of adipose
(2016) 24–35. derived mesenchymal stem cells on synthesized biodegradable Mg–Zn–Ca
[21] Z.B. Tang, J.L. Niu, H. Huang, H. Zhang, J. Pei, J.M. Ou, G.Y. Yuan, Potential alloys, Mater. Sci. Eng., C 69 (2016) 584–597.
biodegradable Zn–Cu binary alloys developed for cardiovascular implant ap- [52] A.J. Drelich, S. Zhao, R.J. GuilloryII, J.W. Drelich, J. Goldman, Long-term surveil-
plications, J. Mech. Behav. Biomed. Mater. 72 (2017) 182–191. lance of Zn implant in murine artery: surprisingly steady biocorrosion rate,
[22] M.A. Arenas, J.D. Damborenea, Protection of Zn–Ti–Cu alloy by cerium trichlo- Acta Biomater. 58 (2017) 539–549.
ride as corrosion inhibitor, Surf. Coat. Technol. 200 (7) (2005) 2085–2091. [53] L. Ponsonnet, K. Reybier, N. Jaffrezic, V. Comte, C. Lagneau, M. Lissac,
[23] G.L. Leone, H.W. Kerr, Grain structures and coupled growth in Zn–Ti alloys, C. Martelet, Relationship between surface properties (roughness, wettability)
J. Cryst. Growth 32 (1) (1976) 111–116. of titanium and titanium alloys and cell behaviour, Mater. Sci. Eng., C 23 (4)
[24] J. Venezuela, M.S. Dargusch, The influence of alloying and fabrication tech- (2003) 551–560.
niques on the mechanical properties, biodegradability and biocompatibility of [54] P. Solař, O. Kylián, A. Marek, M. Vandrovcová, L. Bačáková, J. Hanuš, J. Vyskočil,
zinc: a comprehensive review, Acta Biomater 87 (2019) 1–40. D. Slavínská, H. Biederman, Particles induced surface nanoroughness of tita-
[25] G. Li, H. Yang, Y. Zheng, X.-H. Chen, J.-A. Yang, D. Zhu, L. Ruan, K. Takashima, nium surface and its influence on adhesion of osteoblast-like MG-63 cells,
Challenges in the use of zinc and its alloys as biodegradable metals: perspec- Appl. Surf. Sci. 324 (2015) 99–105.
tive from biomechanical compatibility, Acta Biomater. 97 (2019) 23–45. [55] Q. Li, Z.H. Wang, S. Zhang, W.T. Zheng, Q. Zhao, J. Zhang, L.Y. Wang,
[26] D. Hernández-Escobar, S. Champagne, H. Yilmazer, B. Dikici, C.J. Boehlert, S.F. Wang, D.L. Kong, Functionalization of the surface of electrospun
H. Hermawan, Current status and perspectives of zinc-based absorbable alloys poly(epsilon-caprolactone) mats using zwitterionic poly (carboxybetaine
for biomedical applications, Acta Biomater. 97 (2019) 1–22. methacrylate) and cell-specific peptide for endothelial progenitor cells capture,
[27] J.D. Whitehouse, N.D. Friedman, K.B. Kirkland, W.J. Richardson, D.J. Sexton, The Mater. Sci. Eng., C 33 (2013) 1646–1653.
impact of surgical-site infections following orthopedic surgery at a community [56] E.O. Hall, The deformation and ageing of mild steel: III discussion of results,
hospital and a university hospital adverse quality of life, excess length of stay, Proc. Phys. Soc. B 64 (1951) 747–753.
and extra cost, Infect. Cont. Hosp. Ep. 23 (4) (2002) 183–189. [57] N.J. Petch, The cleavage strength of polycrystals, J. Iron Steel Inst. 174 (1953)
[28] L. Ren, K. Yang, Antibacterial design for metal implants, Met. Foam Bone (2017) 25–28.
203–216. [58] M. Shahin, K. Munir, C.E. Wen, Y.C. Li, Magnesium matrix nanocomposites for
[29] R. Bateni, J.A. Szpunar, X. Wang, D.Y. Li, Wear and corrosion wear of medium orthopedic applications: a review from mechanical, corrosion, and biological
carbon steel and 304 stainless steel, Wear 260 (1) (2006) 116–122. perspectives, Acta Biomater 96 (2019) 1–19.
[30] Z.G. Wang, Y. Li, W.J. Huang, X.L. Chen, H.R. He, Micro-abrasion-corrosion be- [59] P. Vanýsek, Electrochemical Series, CRC Press LLC., 2008.
haviour of a biomedical Ti–25Nb–3Mo–3Zr–2Sn alloy in simulated physiologi- [60] G.Y. Yuan, X.B. Zhang, J.L. Niu, H.R. Tao, D.Y. Chen, Y.H. He, Y. Jiang, W.J. Ding,
cal fluid, J. Mech. Behav. Biomed. Mater. 63 (2016) 361–374. Research progress of new type of degradable biomedical magnesium alloys
[31] X.Y. Yang, C.R. Hutchinson, Corrosion-wear of β -Ti alloy TMZF JDBM, Chin. J. Nonferr. Metal. 21 (10) (2011) 2476–2488.
J. Lin, X. Tong and Z. Shi et al. / Acta Biomaterialia 106 (2020) 410–427 427

[61] T.T. Ren, X. Gao, C. Xu, L.J. Yang, P.S. Guo, H.N. Liu, Y.X. Chen, W.S. Sun, toxicity testing standards for biodegradable magnesium-based materials, Acta
Z.L. Song, Evaluation of as-extruded ternary Zn–Mg–Zr alloys for biomedical Biomater 21 (2015) 237–249.
implantation material: in vitro and in vivo behavior, Mater. Corros. 70 (6) [70] M.S. Dambatta, N.S. Murni, S. Izman, D. Kurniawan, G.R. Froemming, H. Her-
(2019) 1056–1070. mawan, In vitro degradation and cell viability assessment of Zn–3Mg alloy
[62] S. Kumar, M. Chakraborty, V. Subramanya Sarma, B.S. Murty, Tensile and wear for biodegradable bone implants, Proc. Inst. Mech. Eng., Part H 229 (5) (2015)
behaviour of in situ Al–7Si/TiB2 particulate composites, Wear 265 (1–2) (2008) 335–342.
134–142. [71] J. Ma, N. Zhao, D.H. Zhu, Endothelial cellular responses to biodegradable metal
[63] D.B. Liu, B. Wu, X. Wang, M.F. Chen, Corrosion and wear behavior of Zn, ACS Biomater. Sci. Eng 1 (11) (2015) 1174–1182.
an Mg–2Zn–0.2Mn alloy in simulated body fluid, Rare Met. 34 (8) (2015) [72] C. Shen, X.W. Liu, B. Fan, P.H. Lan, F.Y. Zhou, X.K. Li, H.L. Wang, X. Xiao, Li Li,
553–559. S. Zhao, Z. Guo, Z.J. Pu, Y.F. Zheng, Mechanical properties, in vitro degradation
[64] P. Trumbo, A.A. Yates, S. Schlicker, M. Poos, Dietary reference intakes: vita- behavior, hemocompatibility and cytotoxicity evaluation of Zn–1.2Mg alloy for
min A, vitamin K, arsenic, boron, chromium, copper, iodine, iron, manganese, biodegradable implants, RSC Adv. 6 (89) (2016) 86410–86419.
molybdenum, nickel, silicon, vanadium, and Zn, J. Am. Diet. Assoc. 101 (3) [73] L.B. Sukhodub, Antimicrobial activity of Ag+ , Cu2+ , Zn2+ , Mg2+ ions doped chi-
(2001) 294–301. tosan nanoparticles, Anali Mečnikìvsʹkogo Institutu 1 (1) (2015) 39–43.
[65] H. Tapiero, K.D. Tew, Trace elements in human physiology and pathology: zn [74] R. Wahab, A. Mishra, S.I. Yun, Y.S. Kim, H.S. Shin, Antibacterial activity of
and metallothioneins, Biomed. Pharmacother. 57 (2003) 399–411. ZnO nanoparticles prepared via non-hydrolytic solution route, Appl. Microbiol.
[66] S. Gürsoy, A.G. Acar, C. Seşen, Comparison of metal release from new and re- Biotechnol. 87 (5) (2010) 1917–1925.
cycled bracket-archwire combinations, Angle Orthod. 75 (1) (2005) 92–94. [75] G. Grass, C. Rensing, M. Solioz, Metallic copper as an antimicrobial surface,
[67] Y. Oba, A. Yasue, K. Kaneko, R. Uchida, A. Shioyasono, K. Moriyama, Compari- Appl. Environ. Microbiol 77 (5) (2011) 1541–1547.
son of stability of mandibular segments following the sagittal split ramus os- [76] R. Hong, T.Y. Kang, C.A. Michels, Gadura N, Membrane lipid peroxidation in
teotomy with poly-l-lactic acid (PLLA) screws and titanium screws fixation, copper alloy mediated contact killing of Escherichia Coli, Appl. Environ. Micro-
Orthod. Waves 67 (1) (2008) 1–8. biol (2012) 1776–1784.
[68] F. Witte, The history of biodegradable magnesium implants: a review, Acta Bio- [77] H.T. Michels, J.O. Noyce, C.W. Keevil, Effects of temperature and humidity on
mater 6 (2010) 1680–1692. the efficacy of methicillin-resistant Staphylococcus aureus challenged antimi-
[69] J.L. Wang, F. Witte, T.F. Xi, Y.F. Zheng, K. Yang, Y.S. Yang, D.W. Zhao, J. Meng, crobial materials containing silver and copper, Lett. Appl. Microbiol. 49 (2)
Y.D. Li, W.R. Li, K.M. Chan, L. Qin, Recommendation for modifying current cyto- (2009) 191–195.

Das könnte Ihnen auch gefallen