You are on page 1of 11

Research Article

Cite This: ACS Appl. Mater. Interfaces 2019, 11, 6809−6819 www.acsami.org

Mechanical Strength, Biodegradation, and in Vitro and in Vivo


Biocompatibility of Zn Biomaterials
Donghui Zhu,*,† Irsalan Cockerill,† Yingchao Su,† Zhaoxiang Zhang,‡ Jiayin Fu,§ Kee-Won Lee,∥
Jun Ma,† Chuka Okpokwasili,⊥ Liping Tang,⊥ Yufeng Zheng,# Yi-Xian Qin,¶ and Yadong Wang§

Department of Biomedical Engineering, University of North Texas, Denton, Texas 76207, United States

Department of Plastic Surgery and ∥Department of Bioengineering, University of Pittsburgh, Pittsburgh, Pennsylvania 15261,
United States
§
Nancy E and Peter C. Meinig School of Biomedical Engineering, Cornell University, New York 14853, United States
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.


Department of Bioengineering, University of Texas at Arlington, Arlington, Texas 76010, United States
#
Department of Materials Science and Engineering, College of Engineering, Peking University, Beijing 100871, China
Downloaded via UNIV FED DE OURO PRETO on July 25, 2021 at 21:16:29 (UTC).


Department of Biomedical Engineering, Stony Brook University, New York 11794, United States
*
S Supporting Information

ABSTRACT: Zn-based biomaterials have emerged as promising new types of bioresorbable metallics applicable to orthopedic
devices, cardiovascular stents, and other medical applications recently. Compared to other degradable metallic biomaterials (i.e.,
Mg- or Fe-based), Zn biomaterials have a more appropriate corrosion rate without hydrogen gas evolution. Here, we evaluated
the potential of Zn-based metallics as medical implants, both in vitro and in vivo, alongside a standard benchmark Mg alloy,
AZ31. The mechanical properties of the pure Zn were not strong enough but were significantly enhanced (microhardness > 70
kg/mm2, strength > 220 MPa, elongation > 15%) after alloying with Sr or Mg (1.5 at. %), surpassing the minimal design criteria
for load-bearing device applications. The corrosion rate of Zn-based biomaterials was about 0.4 mm/year, significantly slower
than that of AZ31. The measured cell viability and proliferation of three different human primary cells fared better for Zn-based
biomaterials than AZ31 using both direct and indirect culture methods. Platelet adhesion and activation on Zn-based materials
were minimal, significantly less than on AZ31. The hemolysis ratio of red cells (<0.5%) after incubation with Zn-based materials
was also well below the ISO standard of 5%. Moreover, Zn-based biomaterials promoted stem cell differentiation to induce the
extracellular matrix mineralization process. In addition, in vivo animal testing using subcutaneous, bone, and vascular
implantations revealed that the acute toxicity and immune response of Zn-based biomaterials were minimal/moderate,
comparable to that of AZ31. No extensive cell death and foreign body reactions were observed. Taken together, Zn-based
biomaterials may have a great potential as promising candidates for medical implants.
KEYWORDS: Zn alloy, biocompatibility, orthopedic, cardiovascular, corrosion, mechanical property

■ INTRODUCTION
Biodegradable metals are metals that can corrode progressively
metals should be composed of essential metallic elements,
which degrade at desirable rates and can be easily metabolized
in the human body to full dissolution while completing the by the human body through absorption or excretion.1−5
objective to support tissue regeneration without implant Primary bioresorbable metals include iron (Fe), magnesium
residues.1,2 Ideally, the released corrosion products should
induce minimal host response and toxicity, thus initial design Received: November 23, 2018
should avoid using components known to induce these Accepted: January 29, 2019
effects.2 Therefore, the bulk composition of biodegradable Published: January 29, 2019

© 2019 American Chemical Society 6809 DOI: 10.1021/acsami.8b20634


ACS Appl. Mater. Interfaces 2019, 11, 6809−6819
ACS Applied Materials & Interfaces Research Article

(Mg), and zinc (Zn). The past few decades have concentrated and formability of Zn alloys with 1.5% Mg or Sr were found to be
on Fe- and Mg-based biomaterials, but the success of such significantly improved when compared to the pure Zn in our pilot
metals is still slightly limited. Recent studies revealed that Fe studies (data not shown). They were prepared from pure Zn
might be an inappropriate material because of its slower (99.99%), pure Sr (99.99%), or Mg (99.99%) ingots in a high-purity
graphite crucible as described before.22 A gas mixture of 1 vol % SF6/
degradation compared to the healing rate and the long-term
CO2 balance was used to prevent burning and oxidation. After being
retention of a voluminous chemically stable iron oxide.6−8 held at 630 °C for 0.5 h, the melt pool was transferred to a 150 °C
Therefore, Fe can hardly be considered a real “bioresorbable” preheated steel mold. As-cast binary alloys were subsequently hot-
metal. Conversely, Mg degrades too fast, usually within 1−4 rolled or extruded to increase the mechanical properties. Rolled sheets
months, far under the desired time frame of at least 4−6 and the extruded cylinders were then cut into standard testing
months for orthopedic or stent applications.9,10 Also, the specimens for mechanical testing (Vickers hardness and tensile and
significant hydrogen gas evolution and pH change associated uniaxial compression testing) as described before.22 In addition, disk
with Mg degradation in the local environment are another big samples (Φ 10 × 1.5 mm) including AZ31, pure Zn, Zn−Sr, and Zn−
concern in clinical applications. Moreover, current typical Mg Mg alloy specimens were prepared for corrosion tests, hemolysis,
alloys with slightly enhanced corrosion resistance (e.g., AZ- platelet adhesion, and cell experiments. Wires of various diameters
and WE-series alloys) for medical implants usually contain and 10 mm lengths were prepared through extrusion-drawing
methods for implantation into mouse or rat animal models. For all
aluminum, zirconium, and/or rare earth elements, which are
biological testing, specimens were ground up to 4000 grit SiC to
potentially toxic.11−14 While still showing some promises, the smooth out the surface, cleaned ultrasonically for 15 min each in
design of a new Mg-based alloy as an ideal cardiovascular or acetone, ethanol, and distilled water, disinfected with 70% (v/v)
orthopedic implant material presents a significant metallurgical ethanol for 30 min, and UV-sterilized for 2 h.22
challenge. Mechanical Tests. Tension and uniaxial compression tests were
Metallic Zn is a physiologically relevant metal but has not performed according to ASTM-E8M-09 and ASTM E9-89a (2000)
yet been fully considered as a bioresorbable medical implant. standards as described before.22 Vickers hardness was evaluated with a
Compared to titanium and stainless steels, their physical and 100 g applied load and a 15 s loading time. Vickers hardness (Hv) was
mechanical properties are much more in line with human calculated according to the formula: Hv = 1.8544P/d2, where Hv is in
bone.15,16 The typical drawbacks associated with Fe- and Mg- units of kilogram per square millimeter, the indenter load (P) is in
based alloys such as unsuitable corrosion rate and lack of kilogram, and the indenter diagonal length (d) is in millimeter.22
Immersion Tests. According to ASTM-G31, samples were soaked
ductility are usually not applied to Zn-based alloys.17
in simulated body fluid (SBF) solution as described before with slight
Furthermore, Zn2+ is the second most copious element to changes.38,39 At different time intervals, a precalibrated pH meter was
humans, confined primarily to muscles and bones.18−21 used to assess medium pH. No medium was changed or replenished
Importantly, over 600 enzymes rely on Zn2+ for proper during the process of immersion. Then samples were removed from
orientation and function.20 However, research on Zn SBF solution after up to 8 weeks. Specimens were then examined by
biomaterials as medical implants is still limited,22−35 with scanning electron microscopy (SEM) to characterize surface
most of them focused on in vitro analysis of mechanical and morphological changes. To remove corrosion products from the
corrosion properties. surface, specimens were washed with a solution of chromic acid (200
Pure Zn itself is not strong enough for load-bearing medical g/L CrO3 and 10 g/L AgNO3) for a duration of 10 min. Samples were
applications. To improve its mechanical strength, Zn needs to then taken out and dried at room temperature.40,41 Sample weights
be alloyed with other elements to form Zn-based alloys. were determined prior to and after removal of corrosion products.
Although some recent studies showed that alloying with The corrosion rate (C) was calculated according to C = Δm/ρAt,
where C is in units of millimeter per year, Δm is the mass loss, ρ is the
elements including aluminum, rare earth, neodymium, and material density, A is the starting surface area, and t is the time of
yttrium could significantly enhance Zn alloys’ mechanical immersion.
properties,1,2,15 most of these alloying elements are not Cell Culture. Human coronary artery endothelial cells (HCAECs),
desirable because of their potential toxicity. Therefore, group human osteoblast cells (HOBs), and bone marrow-derived human
IIA nutrient elements of the periodic table such as Mg, Ca, and mesenchymal stem cells (hMSCs) were cultured as described
Sr are preferred because they are important mineral supple- before.19,20,38,39,42−46 HCAECs were purchased from ScienCell and
ments for human health.2,36,37 cultured with endothelial cell medium, including 5% fetal bovine
Taken together, these findings inspired us to logically serum (FBS) and supplemented with 1% endothelial cell growth
investigate Zn-based biomaterials as potential alternatives, or supplement and 1% penicillin/streptomycin. HOBs were obtained
even replacements, for medical implants based on their from Thermo Fisher Scientific and cultured in 10% FBS Dulbecco’s
biodegradation, mechanical, and biocompatibility properties. modified Eagle’s medium/F12 supplemented with 1% penicillin/
streptomycin. hMSCs were cultured with nucleotide-free 10% FBS α-
Research studies on this promising biodegradable material are minimum essential medium supplemented with 1% penicillin/
still largely lacking. Using a typical Mg alloy, AZ31, as our streptomycin. Standard culture conditions of 37 °C, 5% CO2, and
comparative benchmark, we examined Zn-based metallics as 95% relative humidity were applied, and cell medium was replaced
potentially better bioresorbable metals for medical applica- every 2−3 days. HCAECs and HOBs at passages 4−6 and hMSCs at
tions, including their mechanical strength, biodegradation, passages 3−5 were used for experiments. ISO 10993-5/12 culture
cytocompatibility, stem cell differentiation, and acute inter- standard was applied to prepare media extracts from the samples.
actions (up to 1 month) with connective, bone, and vascular Cell Viability. Cell viability was evaluated through the 3-[4, 5-
tissues in animal models. dimethylthiazol-2-yl]-2, 5 diphenyltetrazolium bromide (MTT) assay


(Thermo Fisher Scientific, US) as described before.19,20,38,39,42−46
Briefly, cells were seeded into 96-well plates at 10 000 cells/well and
MATERIALS AND METHODS incubated for 24 h to enable attachment. Specific extract with different
Material Preparation. Pure Zn (99.9998%) and extruded concentrations then replaced the cell medium, and cells were cultured
magnesium alloy AZ31 (Mg96/Al3/Zn1) were obtained from for up to 5 days (n = 8). Following this, culture medium with 10%
Goodfellow. Zn binary alloys, Zn−Sr and Zn−Mg with 1.5% (atomic mM MTT solution replaced the extract medium. Sodium dodecyl
%) of alloying elements Sr or Mg, were selected because the plasticity sulfate−HCl (100 μL) was added to the culture wells after 4 h of

6810 DOI: 10.1021/acsami.8b20634


ACS Appl. Mater. Interfaces 2019, 11, 6809−6819
ACS Applied Materials & Interfaces Research Article

incubation time. After another 4 h, absorbance of each well was read implants were placed laterally approximately 15 mm on either side of
at 570 nm. Cells cultured with normal medium were used as control. the incision. Incisions were sealed with surgical clips, and mice were
Cell Proliferation. To evaluate the cytotoxicity of our materials monitored for up to 4 weeks in housing for irritation until explanted.
from different perspectives, the CyQUANT direct cell proliferation In Vivo Bone Implantation. All procedures were approved by
assay (Thermo Fisher Scientific, US) was used to determine the the IACUC at Peking University according to NIH guidelines for the
proliferation rate of different cells as described before.19,20,42 Cells care and use of laboratory animals. We adopted a rat femur condyle
were allowed to attach for 24 h at a density of 10 000 cells/well. Cell model (n = 12 for each different implant) as described before.49
medium was swapped for extract medium, and cells were cultured for Briefly, after being anesthetized, male young Sprague-Dawley rat (2−3
up to 5 days (n = 8). 1× detergent reagent replaced extract solution, months) with an average of 200 g was immobilized, and a longitudinal
and cells were again cultured for 1 h, prior to reading sample plate incision approximately 10 mm in length alongside the lateral extensor
fluorescence (F) from bottom at 480/520 nm. mechanism and close to the knee joint was made. A bone defect,
Cell Adhesion. Cell adhesion analysis was performed as described measuring 5 mm depth and 2 mm in diameter, was created through
before.19,20,39,42,45,46 Briefly, cells were seeded onto the sample the distal bone epiphysis gap. Zn implant was then placed through rat
surfaces, and cells were incubated under standard conditions for 5 h femoral condyles along with sham-operated control. The wound was
(n = 6). Subsequently, culture medium was removed, and the cells closed carefully, and the rats were rested and housed in standard
were rinsed with Dulbecco’s phosphate-buffered saline (DPBS). Cells conditions. The rats were sacrificed after 1 month, and 10% formalin
were stained by Calcein AM prior to fixation using 4% buffer was used to fix femoral condyles for histomorphometric
paraformaldehyde. Adhered cells were visualized by an EVOS FL analysis.
Cell Imaging System fluorescence microscope (AMG, US) and Histomorphometric Analysis of the Bone Tissue. We
further studied by ImageJ. Cell density and retention ratio were processed the explants for bone tissue staining as described before.49
confirmed through a minimum of 10 different fields. Briefly, explants were fixed, ethanol was dehydrated, xylene was
Hemolysis Test. Hemolysis test was performed as described cleared, and methyl methacrylate was embedded. Samples were cut
before.39 Briefly, fresh healthy human blood (Cedarlane Lab, US) was along their vertical axes. After grinding and polishing, section samples
diluted with 1× DPBS at 4:5 volume ratio. Sample pieces (n = 6) were stained with van Gieson Picrofuchsin. Microscopy (Olympus
were placed in 10 mL of 1× DPBS along with 0.2 mL diluted blood. CKX41, Japan) images were utilized to image and examine the
Each tube was then incubated at 37 °C for 1 h before centrifuging at specimens.
3000 rpm. Absorbance of the supernatants was read at 545 nm. In Vivo Vascular Implantation. Cornell University’s IACUC
Positive and negative controls included diluted blood with 10 mL approved all procedures in accordance with NIH guidelines for the
deionized water and 10 mL 1× DPBS, respectively. care and use of laboratory animals. Sprague-Dawley young male rats
Alkaline Phosphatase Activity. The alkaline phosphatase (ALP) (2−3 months, weight = 225−250 g, Charles River Laboratories,
activity assay kit (Thermo Fisher Scientific, US) was used to measure Boston, MA) were used for alloy wire implantation (n = 6 for each
ALP activity by following the manufacturer’s instructions.32 Briefly, different implant). Isoflurane induction (5%) and 2% maintenance
stem cells with a seeding density of 6000 cells/cm2 were cultured in a were used to anesthetize the rats. An abdominal midline incision 20
cell differentiation medium atop material surfaces for a total of 7 days mm in length exposed the abdominal aorta, which was then separated
with medium changed every 3 days (n = 8). Then, cell lysate was from the inferior vena cavas. Both proximal and distal parts of the
collected, and total protein was determined by the bicinchoninic acid aorta were clamped with microvascular movable clips. Wires made of
assay. Next, supernatants were mixed with a StemTAG AP Activity AZ31, Zn, Zn−Sr, and Zn−Mg in the size of 0.25 mm diameter were
Assay Substrate for 30 min at 37 °C. Addition of 1× Stop Solution cut as 10 mm length, rinsed with acetone, and disinfected for 30 min
completed the reaction, and absorbance was read at 405 nm, with 70% (v/v) ethanol and 2 h with UV prior to implantation. To
normalized by the total protein, which was used to determine the ALP contact the wire with circulating arterial blood, both ends were
activity. inserted into the lumen throughout the proximal and distal parts of
Alizarin Red Activity. hMSC osteogenic differentiation was the arterial wall by puncturing the arterial adventitia and leading the
assessed via the Alizarin Red quantification assay by following the wire in, followed by a second puncturing in the arterial
manufacturer’s instructions (Thermo Fisher Scientific, US).32 Briefly, adventitia.50−52 Microvascular movable clips were removed from the
stem cells were cultured for 20 days on material surfaces in plates (n = aorta, and the instant blood flow was checked with a flow Doppler for
8). Cells were lysed from the material surface using 10% acetic acid, 5, 10, and 30 min. The surgical site was closed with 5−0 Dacron
and lysates were heated for 10 min at 85 °C. Cell lysates were then sutures. No anticoagulation or antiplatelet treatments were adminis-
cooled for 5 min on ice, prior to 15 min centrifugation at 20 000g. trated pre- and postoperatively. Antibiotic (Baytril, 20 mg/kg) and
Then, 200 μL supernatant was collected and mixed with the same analgesic (Buprenex, 0.02 mg/kg) medications were given until 72 h
volume of ammonium hydroxide to neutralize the acid. ARS postoperatively.
concentration was obtained from the supernatant absorbance at 405 Vascular Explantation and Histological Analysis. At 1 month
nm and a preprepared standard curve. postimplantation, rats were anesthetized, and a midline incision
RT-PCR for the mRNA Measurement. Collagen I and exposed the abdominal aorta. The aorta including Zn wire was taken
osteopontin gene expression level also indicate the osteodifferentia- out from surrounding tissues for analysis. The explanted wire/aorta
tion of the hMSCs, which is examined by CFX96 Touch RT-PCR was rinsed with sterile phosphate-buffered saline, fixed for 1 h in 10%
Detection System (Bio-Rad, US) as described before.32 Total RNA neutral formalin, and soaked for 24 h at 4 °C in 30% sucrose solution.
was extracted [RNeasy Mini Kit (Qiagen, US)] and calculated with a The whole wire/aorta was embedded vertically into the optimal
Nanodrop spectrophotometer. RNA was reverse-transcribed in a cutting temperature (OCT) compound (Sakura Finetek, Torrance,
thermo cycler (T100, Bio-Rad, US) using the RT2 First Strand Kit CA), snap-frozen at −80 °C, and cryosectioned as 10 μm thickness.
(Qiagen, US). β-Actin mRNA level served as the endogenous control Hematoxylin and eosin (H&E) staining of the slides was done on the
and reference. Final mRNA expression data of target protein were vascular tissue surrounding the wire.
obtained using the 2 − ΔΔCt method with Bio-Rad CFX Manager 3.1 Hematoxylin and Eosin. At 1 month postimplantation,
software. surrounding tissues including implants were excised for histological
In Vivo Subcutaneous Implantation. The University of Texas evaluation and immunofluorescence staining. Specifically, biomaterial-
at Arlington’s IACUC approved all procedures in accordance with induced inflammatory reactions were measured by quantifying
NIH care and use of laboratory animal guidelines. All metallic macrophages, neutrophils, mast cells, and fibrocytes at the implant−
implants (AZ31, Zn, Zn−Sr, and Zn−Mg) and sham-operated control tissue interface as described in our works.47,53−57 Surrounding tissues
received dorsal midline incisions following anesthetization in C57 and alloys were excised for frozen sectioning by embedding in OCT.
mice (Jackson Labs; n = 6) as previously described.47,48 Briefly, a Then, cross sections (7 μm) were made, and cell infiltration of the
subcutaneous incision was made in each mouse, and two metallic different tissue layers was visualized by H&E staining. Interface

6811 DOI: 10.1021/acsami.8b20634


ACS Appl. Mater. Interfaces 2019, 11, 6809−6819
ACS Applied Materials & Interfaces Research Article

Figure 1. Mechanical properties of Zn, AZ31, Zn−Sr, and Zn−Mg alloys: (A) microhardness; (B) tensile properties; and (C) elongation. Scale bar
= 2.5 mm (aP < 0.05, bP < 0.01, cP < 0.001).

Figure 2. Immersion test of AZ31 and Zn materials in a time course up to 30 days. (A) Pictures of AZ31 (left) and Zn (right) after up to 30 days of
immersion in SBF; (B) cumulative Mg ion released from AZ31; (C) cumulative Zn ion released from Zn; (D) change of pH value in solution; and
(E) calculated corrosion rate of Zn and AZ31 from immersion testing (cP < 0.001).

response quantification was carried out using ImageJ as reported Corrosion Behavior. Ideally, the corrosion or biodegra-
previously.47,48 Briefly, images taken on the skin side of the tissue− dation rate of implants should meet the pace of tissue healing.
biomaterial interface are shown as the multiple count averages of We then examined the corrosion behavior of the materials
H&E stains.
based on immersion testing. Figure 2A shows the surface
Statistical Analysis. All data were presented as mean ± standard
deviation. Assays were repeated 3 times autonomously with a morphology of samples soaked in SBF up to 30 days. The Zn
minimum of three replicates. One-way or two-way analysis of samples retained flat surfaces after 30 days, whereas the
variance followed by Tukey’s post-hoc test or two-tailed Student’s t- surfaces of AZ31 were covered with corrosion products.
test, as appropriate, was used to analyze the statistical significance. P < Cumulative Mg ion released from the alloy AZ31 increased
0.05 was considered as statistically significant. linearly over the immersion period, whereas cumulative Zn ion

■ RESULTS
Mechanical Strength. Mechanical properties are signifi-
released from Zn materials reached a plateau after ∼20 days,
indicating slower corrosion after initial stage for Zn (Figures
2B,C and S1). Immersion medium pH change was recorded
cant factors for load-bearing bioresorbable metals (e.g., bone (Figures 2D and S1). Degradation of AZ31 induced significant
plates/screws and vascular stents); thus, they should have pH increase compared to control and Zn over time. However,
enough strength comparable to the local tissues as for degradation of Zn did not induce significant pH change. Figure
mechanical performance. The mechanical properties of pure 2E shows the calculated corrosion rate of AZ31 and Zn-based
Zn are poor. To meet the needs of the implant, strengthening materials from immersion testing. The corrosion rate of Zn-
strategies include alloying and hot-working deformation, which based materials was approximately 0.4 mm/year significantly
were tested here.39 After alloying with Sr or Mg, the slower than that of AZ31.
microhardness of Zn increased from 40 to over 70 kg/mm2 Cytocompatibility by Indirect Culture. Primary
(Figure 1A). The yield strength (YS) and ultimate tensile HCAECs, HOBs, and hMSCs were cultured with a
strength (UTS) were significantly improved from less than 50 concentration gradient of extract solutions from Zn and
to over 220 MPa after alloying (Figure 1B). A similar trend was AZ31 materials over the course of 1 (Figure 3), 3 (Figure S2),
observed for elongation, which increased from ∼5 to over 15% and 5 days (Figure S3). Increasing extract concentration
after alloying with Sr or Mg (Figure 1C). It is noteworthy that reduced HCAEC viability and proliferation in an overall trend
Zn−Sr and Zn−Mg had higher microhardness and elongation (Figure 3A,B). However, cell viability and proliferation
than that of benchmark control, AZ31, while they all had reduced much more significantly for AZ31 than Zn at higher
similar YS and UTS. Such data demonstrate that through concentrations (P < 0.05). This is probably due to the
alloying, Zn-based biomaterials can be developed with increasing levels of ions (i.e., Mg2+, Al3+, and Zn2+) and pH in
sufficient strength, rendering them as appropriate load-bearing the extract solution of AZ31. Moreover, different cells have
biomedical implants. different sensitivities or tolerances to extracellular ion
6812 DOI: 10.1021/acsami.8b20634
ACS Appl. Mater. Interfaces 2019, 11, 6809−6819
ACS Applied Materials & Interfaces Research Article

Figure 4. Direct culture of primary HCAECs, HOBs, and hMSCs on


the surfaces of AZ31 and Zn plates for 1−5 days with quantitative cell
adhesion density measurements: (A) cell attachment on AZ31; (B)
quantitative measurement of cell density on AZ31; (C) cell
attachment on Zn; and (D) quantitative measurement of cell density
on Zn (aP < 0.05).

Figure 3. Culture of primary HCAECs, HOBs, and hMSCs with


extracts of AZ31 and Zn materials for 1 day. Cell viability (A,C,E) and
cell proliferation (B,D,F) were measured separately for these three cell
types, respectively (aP < 0.05, bP < 0.01, cP < 0.001).

concentrations.19,20,42 Interestingly, the cell viability and


proliferation of HOBs and hMSCs did not decrease after
treatment with Zn or AZ31 medium extracts (Figure 3C−F).
When the extract concentrations were lower, they promoted
the cell growth of HOBs and hMSCs (P < 0.05). Among the
three cell types, HCAECs were the most sensitive to the
medium extracts and had the least tolerance. AZ31 was the
most toxic to HCAECs as most HCAECs died after 3−5 days
of culture. On the contrary, no significant cell toxicity of AZ31
and Zn was observed for HOBs and hMSCs at the same
culture conditions (Figures 3, S2, and S3).
Cytocompatibility by Direct Culture. We also per-
formed in vitro cytocompatibility testing by direct culture of
primary HCAECs, HOBs, and hMSCs on surfaces of Zn and
AZ31 plates for up to 5 days. For HCAECs, there were hardly
any viable cells (green) on AZ31, whereas there were
numerous viable cells on Zn (Figure 4A,C). The density of
HCAECs on Zn increased significantly with increasing culture
time, indicating fast growth and proliferation on the Zn surface
(Figure 4B,D). For both HOBs and hMSCs, confluent cell
densities on Zn were achieved over 5 day time course (Figure
4B,D). Cell densities on AZ31 were initially low, but over time Figure 5. Hemocompatibility analysis of AZ31 and Zn materials. (A)
SEM images of platelet adhesion on the sample surface; (B)
they increased significantly, reaching subconfluency after 5
quantitative amount of the adhered platelet; and (C) hemolysis of
days (Figure 4A,C). Similar cytocompatibility results were AZ31 and Zn-based materials (aP < 0.05, bP < 0.01).
observed for Zn−Sr and Zn−Mg materials (Figure S4).
Hemocompatibility. We carried out the blood compati-
bility testing as shown in Figure 5. The adhered platelet mostly of round shape, lacking pseudopodia spreading (Figures
morphology on plates of AZ31 and Zn-based metallics was 5A and S5), indicating no/minimal platelet activation. The
6813 DOI: 10.1021/acsami.8b20634
ACS Appl. Mater. Interfaces 2019, 11, 6809−6819
ACS Applied Materials & Interfaces Research Article

number of platelets on the top of Zn-based materials was far found that tissue responses to Zn-based materials were
less than on AZ31 (Figure 5B). These results indicate superb comparable to those of AZ31 (Figures 8 and S6). Only a
antithrombotic and antiplatelet adhesion properties. Our
studies also showed relatively low (<0.4%) (Figure 5C)
hemolysis rates on Zn-based materials, which are far below
ISO 10993-4:2002 5% safe value standards, indicating that Zn
alloys will not cause serious hemolysis.
In Vitro Mineralization. It has been reported that Zn
materials promote bone regeneration in vivo.49 Thus, we
examined the in vitro mineralization process at the cell−
material interface by ALP and Alizarin Red staining.
Interestingly, Zn-based materials stimulated significantly
more ALP activity of hMSCs as compared to controls when
the differentiation medium was applied (Figure 6A). In

Figure 6. Osteogenic differentiation of hMSCs as measured by ALP


(A) and Alizarin Red (B) with the treatment of extracts of AZ31 and
Zn-based materials (10% as referred to Figure 3) (aP < 0.05, bP < Figure 8. H&E staining of tissues surrounding different implants for 4
0.01). weeks. (A) Representative images of H&E and (B) quantitative cell
numbers around the implants (* implantation area).
addition, Zn-based materials induced significantly more calcific
deposition and activity of hMSCs evidenced by the Alizarin
Red level staining by the levels of Alizarin Red staining small number of inflammatory cells were found at the interface
compared to controls (Figure 6B). Moreover, osteopontin and between the tissue and alloys. The population of macrophages
collagen I mRNA levels were significantly higher as compared around the Zn-based implants was also comparable to that of
to control, and Zn-based materials had even slightly more AZ31, as measured by CD11b staining (Figures 9 and S7).
mRNA expression of these genes compared to AZ31 (Figure
7A,B), consistent with a preceding study.32

Figure 7. Osteogenic differentiation of hMSCs after different


treatments as measured by mRNA levels of collagen I (Col I) and
osteopontin (OPN). (A) Fold change of the Col I mRNA level as
compared to control. (B) Fold change of the OPN mRNA level as
compared to control (aP < 0.05).

In Vivo Immune Response. We first evaluated the in vivo Figure 9. CD11b staining of the tissues surrounding different
tissue biocompatibility of our Zn biomaterials and AZ31 implants after 2 weeks. (A) Representative images of CD11b positive
control using a subcutaneous implantation model in mice. The cells and (B) quantitative cell numbers around the implants. The red
extent of tissue responses to the alloy implant was assessed arrowhead points to a representative CD11b positive cell, and the
based on the extent of inflammatory responses and fibrotic green arrowhead points to a representative CD11b negative cell (*
tissue formation at the implant−host tissue interface. We implantation area).

6814 DOI: 10.1021/acsami.8b20634


ACS Appl. Mater. Interfaces 2019, 11, 6809−6819
ACS Applied Materials & Interfaces Research Article

Figure 10. Histological characterization of the vascular tissue in a rat abdominal aorta implantation model. (A) Explanted pure Zn wire/aorta at
day 30 postimplantation. (B) Top: H&E staining of the complete cross section in different regions and bottom: partial magnification of the box
shown in top images. (C) H&E staining of the healthy aorta control. Scale bar = 200 μm (w: wire).

In Vivo Vasculature−Implant Interaction. Next, we biomaterials after alloying generally have good mechanical
assessed in vivo vascular cell/tissue responses of pure Zn wires strength suitable for load-bearing implants such as bone
by implanting them in rat abdominal aorta for 1 month (Figure fixation implants and cardiovascular stents, better corrosion
S8). In this study, the explanted tissues appeared to be normal rates matching to the tissue healing pace, and good
and comparable to sham-operated control, indicating no cytocompatibility and tissue compatibility with minimal
embolism and thrombosis in the animals tested (Figure toxicity and immune responses in animal models.
10A). H&E staining of the explants showed the foreign Alloying with Sr or Mg Enhances the Mechanical
response to the wires. There were some inflammatory cells and Strength of Zn with Minimal Effect on Degradation and
tissues in the adventitial side, which underwent fibrotic Biocompatibility. As expected, alloying with Mg or Sr
responses, resulting in layers of infiltrated inflammatory cells elements can significantly enhance the mechanical properties
and extracellular matrix (ECM) deposition (Figure 10B). No of Zn materials to surpass the minimal design requirements of
extensive inflammation was observed. Same experiments were
load-bearing medical implants, consistent with previous
performed for AZ31 and Zn alloys, and similar results were
studies.23,26,49 Mechanical characteristics of some Zn and
observed (data not shown).
In Vivo Bone−Implant Interaction. In addition, we various Zn alloys in comparison to that of Mg and natural bone
tested if Zn implant will promote osteointegration, osteocon- are summarized in Table 1. Similar to the Mg and Mg alloys,
duction, and/or osteoinduction using a rat femoral condyle alloying increased the mechanical properties of Zn, including
implantation model. New bone formation was observed mechanical strength and elongation. The mechanical strength
surrounding the Zn implant via van Gieson staining with an of two Zn alloys in this study is comparable to that of AZ31,
abundance of osteocytes in the newly formed bone (Figure while their elongations are higher than that of the AZ31 alloy.
11). There were also some fibrotic and collagenous tissues Although Zn−Li alloys seem to have the strongest strength,
our results are still in line with previous studies.22,59
Pure metals usually have lower strength compared to those
alloyed with other metals, so does Zn. After alloying, the
formation of intermetallics and different metallic phases is
thought to reduce the movement of dislocations within the
metallic structure. The reasons that Sr or Mg alloying can make
Zn stronger are that two basically structural changes may
Figure 11. Histological characterization of bone tissue sections at happen. The first is solid solution strengthening that involves
implant sites with Van Gieson staining 1 month postimplantation of the addition of other metallic elements dissolving in the parent
Zn. The yellow rectangles correspond to the magnified bone−implant lattice and causing distortions because of the difference in
interface. The green triangle indicates a newly formed bone (* atom size between the parent metal and the solute metal. In
implantation area).
addition, adding elements that have no or partial solubility in
the parent metal will result in the appearance of a second phase
between the implants and newly formed bones. Moreover,
stretching outward from the biomaterial, expanding bone mass distributed throughout the crystal or between crystals, which
was observed, which indicated an osteogenic property of Zn. may raise or reduce the strength of an alloy. In this study, a
Similar results were observed for Zn−Sr and Zn−Mg alloys trace amount of Sr or Mg was used to avoid potential overdose
(data not shown), consistent with previous studies.22,58 toxicity despite their natural existence in human body as a


supplemental nutrition. Therefore, attention should be paid to
DISCUSSION the selection of the alloying elements. Preference should be
Biodegradable metals are designed to provide a temporary given to nutrient or, at least, nontoxic elements to avoid
support after implantation, with the expectation of being potential add-on toxicity from the alloying additions. Although
completely reabsorbed during the tissue healing process. There significant mechanical strengthening was observed after
are three main biodegradable metals Mg, Zn, and Fe. Zn alloying, there were no obvious changes in corrosion rate
biodegradable metals emerge as a new generation of metallic and biocompatibility compared to pure Zn. Data also showed
implants largely for orthopedic and cardiovascular stent that there was no significant difference between Zn−Sr and
applications lately. Using a typical Mg alloy, AZ31, as a Zn−Mg for all the mechanical strength, corrosion, and toxicity
comparative standard control, we demonstrated that Zn tested so far.
6815 DOI: 10.1021/acsami.8b20634
ACS Appl. Mater. Interfaces 2019, 11, 6809−6819
ACS Applied Materials & Interfaces Research Article

Table 1. Summary of Mechanical Properties of Various Implant Materials and Natural Bone
materials UTS (MPa) YS (MPa) elongation (%) compressive YS (MPa)
natural bone 100−15063 1−363 130−1809
pure Mg 85 (as-cast) 20 (as-cast) 10−163 65−1009
165 (hot-rolled)3 110 (hot-rolled)3
pure Zn49,64 18 (as-cast) 10 (as-cast) 0.2 (as-cast) 105 (extruded)
50 (hot-rolled) 30 (hot-rolled) 5.5 (hot-rolled)
60−110 (extruded) 33−50 (extruded) 3.5−60 (extruded)
Zn1X (Mg, Ca, Sr)31,49,64,65 150−185 (as-cast) 120−130 (as-cast) 1−2 (as-cast) 280−340 (extruded)
220−250 (hot-rolled) 188−205 (hot-rolled) 12−20 (hot-rolled)
240−270 (extruded) 200−220 (extruded) 8−11 (extruded)
Zn3Mg31,49,64,65 30 (as-cast) 291 (extruded) 0.2−0.4 (as-cast)
399 (extruded) 1 (extruded)
Zn4Li34 450 (hot-rolled) 420 (hot-rolled) 14 (hot-rolled)
Zn(1−3)Al35,64 220−240 (hot-rolled) 134−210 (hot-rolled) 22−32 (hot-rolled)
AZ31a 280 245 13
pure Zna 50 30 6
Zn-1.5Mga 285 240 17
Zn-1.5Sra 250 220 22
a
Data from this study.

Zn-Based Metallics Have a More Suitable Degrada- In Vitro and in Vivo Biocompatibility of Zn-Based
tion Rate Than That of Mg and Fe. The major problems Implants Is Generally Good. Biocompatibility is another
with Mg-based metals are the too fast degradation rate, the important factor for degradable biomaterials besides mechan-
hydrogen gas evolution, and insufficient mechanical strength ical and corrosion properties. Overall, Zn materials exhibit
for some load-bearing applications. The major drawbacks with comparable, sometimes better hemocompatibility and cyto-
Fe-based metals are the prolonged degradation time and the compatibility than the comparative control AZ31. This could
hard-to-clear degradation products retained in the local tissues. be due to its slow degradation rate, less local pH change, and
The degradation rate of Zn is between that of Mg and Fe minimal hydrogen gas evolution. Testing conditions, param-
because of standard electrode potential (−0.76 V/SCE) of Zn eters, and cell types could be other factors as well. In animal
falling in the middle of Mg (−2.37 V/SCE) and Fe (−0.44 V/
models, the toxicity and immune responses of Zn materials
SCE).15,31 This explains why we observed a moderate and
were also minimal and comparable to AZ31. It is notable that
similar degradation rate of pure Zn and two Zn alloys. It might
also be due to the formation of similar passive degradation Zn alloys promoted significant bone regeneration in a mouse
films for all Zn-based metallics. model as reported recently,49 consistent with our in vitro
Systemic Toxicity of Zn Implants Is Minimal. Similar to mineralization and in vivo data. Formation of a new bone was
Mg or Fe, Zn is an essential trace element. Zn deficiency may visualized by histological analysis at week 4. Some fibrotic and
cause significant disorders, and implant-derived Zn ions collagenous tissues between the new bone tissue and materials
released inside the body may be beneficial if not overdosed.60 were also observed. This could be explained by the high Zn ion
Naturally, overdose is always a concern as Zn’s tolerable upper concentration in the vicinity of implants, which may hinder the
intake level (UL) is 15−40 mg/day, significantly lower than formation of a new bone. Therefore, fibrotic tissues may form
that of Mg (300−400 mg).61 Thanks to its much slower during bone tissue healing to cause a delayed osseointegration.
corrosion rate than that of Mg, the accumulated Zn ion inside Zn degradation products include zinc oxide, zinc hydroxide,
the body is anticipated to be below the UL value. Therefore, and Ca/P compounds, in addition to zinc and hydroxide ions.
the risk of systemic Zn toxicity associated with Zn alloys is As an essential element for various biological functions, Zn
negligible based on the following observations.2 Here is our may promote new bone formation as demonstrated by
rationale. Assuming that a typical Zn implant weighs ∼250 mg enhanced mineralization of the ECM, differentiation of
of pure metal with a time frame of 12 months for a total hMSCs, and enhanced expression of genes related to bone
degradation, the expected daily dose of Zn will be <1 mg/day.
remodelings such as ALP, collagen I, and osteopontin. Zn may
As the recommended daily value of Zn is about 10 mg/day,2
also inhibit osteoclast differentiation. On the other hand,
the amount of Zn released from the degradation of the Zn
extreme Zn ion release locally may induce adverse effects on
alloy implant is only 10% of the recommended daily intake for
adults, which is far below its recommended daily intake value. cells, leading to potential bone resorption in vivo. Therefore,
Thus, Zn implant systemic toxicity appears not a concern, even large concentration of Zn ions released from rapidly
with multiple devices placed in the same patient. The local degradable implants should be carefully controlled.
toxicity should also be minimal as demonstrated by the data Our studies here generally focused on in vitro and acute in
from previous studies22 and our studies here, possibly due to vivo evaluations, and a long-term experiment is needed to fully
the swift transport and clearance of metallic ions, including the understand the potential of Zn-based biomaterials for different
Zn ion, from the tissue to peripheral blood flow,19,20 thus biomedical applications. Nonetheless, there is no doubt that
prohibiting its enrichment and cytotoxicity in the vicinity of Zn-based biomaterials may have a greater potential for medical
the implant. implants.62
6816 DOI: 10.1021/acsami.8b20634
ACS Appl. Mater. Interfaces 2019, 11, 6809−6819
ACS Applied Materials & Interfaces Research Article

■ CONCLUSIONS
Using a typical Mg alloy, AZ31, as a comparative benchmark
(5) Zhou, W. R.; Zheng, Y. F.; Leeflang, M. A.; Zhou, J. Mechanical
property, biocorrosion and in vitro biocompatibility evaluations of
Mg-Li-(Al)-(RE) alloys for future cardiovascular stent application.
control, we examined the potential of some Zn biomaterials as Acta Biomater. 2013, 9, 8488−8498.
medical implants, both in vitro and in vivo. We found that Zn (6) Bowen, P. K.; Gelbaugh, J. A.; Mercier, P. J.; Goldman, J.;
biomaterials after alloying exhibited sufficient mechanical Drelich, J. Tensile testing as a novel method for quantitatively
strength for load-bearing device applications. The corrosion evaluating bioabsorbable material degradation. J. Biomed. Mater. Res. B
rate of Zn biomaterials is significantly slower than that of 2012, 100, 2101−2113.
AZ31. Degradation of AZ31 also induced significant pH (7) Bowen, P. K.; Drelich, J.; Buxbaum, R. E.; Rajachar, R. M.;
increase over time but not for Zn-based materials. The Goldman, J. New approaches in evaluating metallic candidates for
measured blood compatibility, cell viability, and proliferation bioabsorbable stents. Emerg. Mater. Res. 2012, 1, 237−255.
of three different human primary cells fared better for Zn (8) Pierson, D.; Edick, J.; Tauscher, A.; Pokorney, E.; Bowen, P.;
Gelbaugh, J.; Stinson, J.; Getty, H.; Lee, C. H.; Drelich, J.; Goldman,
biomaterials than AZ31. Zn biomaterials also promoted stem
J. A simplified in vivo approach for evaluating the bioabsorbable
cell differentiation to induce the ECM mineralization process. behavior of candidate stent materials. J. Biomed. Mater. Res. B 2011,
Moreover, in vivo animal testing revealed that the toxicity and 100, 58−67.
immune response of Zn biomaterials were minimal/moderate, (9) Staiger, M. P.; Pietak, A. M.; Huadmai, J.; Dias, G. Magnesium
comparable to that of AZ31. No extensive cell death and and its alloys as orthopedic biomaterials: A review. Biomaterials 2006,
foreign body reactions were observed. Thus, Zn biomaterials 27, 1728−1734.
are demonstrated to be promising bioresorbable metal (10) Witte, F.; Hort, N.; Vogt, C.; Cohen, S.; Kainer, K. U.;
contenders for cardiovascular and orthopedic applications. Willumeit, R.; Feyerabend, F. Degradable biomaterials based on


magnesium corrosion. Curr. Opin. Solid St. M. 2008, 12, 63−72.
ASSOCIATED CONTENT (11) Gu, X. N.; Zhou, W. R.; Zheng, Y. F.; Cheng, Y.; Wei, S. C.;
Zhong, S. P.; Xi, T. F.; Chen, L. J. Corrosion fatigue behaviors of two
*
S Supporting Information
biomedical Mg alloys - AZ91D and WE43 - In simulated body fluid.
The Supporting Information is available free of charge on the Acta Biomater. 2010, 6, 4605−4613.
ACS Publications website at DOI: 10.1021/acsami.8b20634. (12) Li, M.; Cheng, Y.; Zheng, Y. F.; Zhang, X.; Xi, T. F.; Wei, S. C.
Immersion tests of Zn−Sr and Zn−Mg materials; cell Surface characteristics and corrosion behaviour of WE43 magnesium
alloy coated by SiC film. Appl. Surf. Sci. 2012, 258, 3074−3081.
viability and proliferation of primary HCAECs, HOBs, (13) Ye, C. H.; Zheng, Y. F.; Wang, S. Q.; Xi, T. F.; Li, Y. D. In vitro
and hMSCs with extracts of AZ31 and Zn materials for 3 corrosion and biocompatibility study of phytic acid modified WE43
and 5 days, respectively; interactions between cells and magnesium alloy. Appl. Surf. Sci. 2012, 258, 3420−3427.
Zn−Mg or Zn−Sr; hemocompatibility analysis of Zn−Sr (14) Ye, C.-h.; Xi, T.-f.; Zheng, Y.-f.; Wang, S.-q.; Li, Y.-d. In vitro
and Zn−Mg; H&E staining of tissues surrounding Zn− corrosion and biocompatibility of phosphating modified WE43
Sr and Zn−Mg; CD11b staining of the tissues magnesium alloy. T. Nonferr. Metal. Soc. 2013, 23, 996−1001.
surrounding Zn−Sr and Zn−Mg implants; and (15) Bowen, P. K.; Drelich, J.; Goldman, J. Zinc Exhibits Ideal
implantation of the Zn wire in the rat abdominal aorta Physiological Corrosion Behavior for Bioabsorbable Stents. Adv.
(PDF) Mater. 2013, 25, 2577−2582.


(16) Bowen, P. K.; Guillory, R. J.; Shearier, E. R.; Seitz, J.-M.;
Drelich, J.; Bocks, M.; Zhao, F.; Goldman, J. Metallic zinc exhibits
AUTHOR INFORMATION optimal biocompatibility for bioabsorbable endovascular stents. Mat.
Corresponding Author Sci. Eng. C Mater. Biol. Appl. 2015, 56, 467−472.
*E-mail: Donghui.Zhu@unt.edu. Phone: 940-369-8707. (17) Drelich, A. J.; Zhao, S.; Guillory, R. J., 2nd; Drelich, J. W.;
Goldman, J. Long-term surveillance of zinc implant in murine artery:
ORCID Surprisingly steady biocorrosion rate. Acta Biomater. 2017, 58, 539−
Donghui Zhu: 0000-0002-3057-1889 549.
Yufeng Zheng: 0000-0002-7402-9979 (18) Guillory, R. J.; Bowen, P. K.; Hopkins, S. P.; Shearier, E. R.;
Earley, E. J.; Gillette, A. A.; Aghion, E.; Bocks, M.; Drelich, J. W.;
Notes Goldman, J. Corrosion Characteristics Dictate the Long-Term
The authors declare no competing financial interest.


Inflammatory Profile of Degradable Zinc Arterial Implants. ACS
Biomater. Sci. Eng. 2016, 2, 2355−2364.
ACKNOWLEDGMENTS (19) Ma, J.; Zhao, N.; Zhu, D. Endothelial Cellular Responses to
This work was supported by the National Institutes of Health Biodegradable Metal Zinc. ACS Biomater. Sci. Eng. 2015, 1, 1174−
(grant number R01HL140562). The content is solely the 1182.
(20) Ma, J.; Zhao, N.; Zhu, D. Bioabsorbable zinc ion induced
responsibility of the authors and does not necessarily represent
biphasic cellular responses in vascular smooth muscle cells. Sci. Rep.
the official views of the National Institutes of Health.


2016, 6, 26661.
(21) Shearier, E. R.; Bowen, P. K.; He, W.; Drelich, A.; Drelich, J.;
REFERENCES Goldman, J.; Zhao, F. In Vitro Cytotoxicity, Adhesion, and
(1) Li, H.; Zheng, Y.; Qin, L. Progress of biodegradable metals. Prog. Proliferation of Human Vascular Cells Exposed to Zinc. ACS
Nat. Sci-Mater. 2014, 24, 414−422. Biomater. Sci. Eng. 2016, 2, 634−642.
(2) Zheng, Y. F.; Gu, X. N.; Witte, F. Biodegradable metals. Mat. Sci. (22) Li, H. F.; Xie, X. H.; Zheng, Y. F.; Cong, Y.; Zhou, F. Y.; Qiu,
Eng. R 2014, 77, 1−34. K. J.; Wang, X.; Chen, S. H.; Huang, L.; Tian, L.; Qin, L.
(3) Gu, X.; Zheng, Y.; Cheng, Y.; Zhong, S.; Xi, T. In vitro corrosion Development of biodegradable Zn-1X binary alloys with nutrient
and biocompatibility of binary magnesium alloys. Biomaterials 2009, alloying elements Mg, Ca and Sr. Sci. Rep. 2015, 5, 10719.
30, 484−498. (23) Li, H.; Yang, H.; Zheng, Y.; Zhou, F.; Qiu, K.; Wang, X. Design
(4) Li, Z.; Gu, X.; Lou, S.; Zheng, Y. The development of binary Mg- and characterizations of novel biodegradable ternary Zn-based alloys
Ca alloys for use as biodegradable materials within bone. Biomaterials with IIA nutrient alloying elements Mg, Ca and Sr. Mater. Design
2008, 29, 1329−1344. 2015, 83, 95−102.

6817 DOI: 10.1021/acsami.8b20634


ACS Appl. Mater. Interfaces 2019, 11, 6809−6819
ACS Applied Materials & Interfaces Research Article

(24) Liu, X.; Sun, J.; Qiu, K.; Yang, Y.; Pu, Z.; Li, L.; Zheng, Y. (41) Ma, J.; Zhao, N.; Zhu, D. Sirolimus-eluting dextran and
Effects of alloying elements (Ca and Sr) on microstructure, polyglutamic acid hybrid coatings on AZ31 for stent applications. J.
mechanical property and in vitro corrosion behavior of biodegradable Biomater. Appl. 2015, 30, 579−588.
Zn-1.5Mg alloy. J. Alloy. Compd. 2016, 664, 444−452. (42) Ma, J.; Zhao, N.; Zhu, D. Biphasic responses of human vascular
(25) Liu, X.; Sun, J.; Yang, Y.; Pu, Z.; Zheng, Y. In vitro investigation smooth muscle cells to magnesium ion. J. Biomed. Mater. Res. A 2015,
of ultra-pure Zn and its mini-tube as potential bioabsorbable stent 104, 347−356.
material. Mater. Lett. 2015, 161, 53−56. (43) Ma, J.; Zhao, N.; Betts, L.; Zhu, D. Bio-Adaption between
(26) Liu, X.; Sun, J.; Yang, Y.; Zhou, F.; Pu, Z.; Li, L.; Zheng, Y. Magnesium Alloy Stent and the Blood Vessel: A Review. J. Mater. Sci.
Microstructure, mechanical properties, in vitro degradation behavior Technol. 2016, 32, 815−826.
and hemocompatibility of novel Zn-Mg-Sr alloys as biodegradable (44) Zhao, N.; Zhu, D. Endothelial responses of magnesium and
metals. Mater. Lett. 2016, 162, 242−245. other alloying elements in magnesium-based stent materials. Metal-
(27) Liu, X.; Sun, J.; Zhou, F.; Yang, Y.; Chang, R.; Qiu, K.; Pu, Z.; lomics 2015, 7, 118−128.
Li, L.; Zheng, Y. Micro-alloying with Mn in Zn-Mg alloy for future (45) Zhao, N.; Zhu, D. Collagen self-assembly on orthopedic
biodegradable metals application. Mater. Design 2016, 94, 95−104. magnesium biomaterials surface and subsequent bone cell attachment.
(28) Shen, C.; Liu, X.; Fan, B.; Lan, P.; Zhou, F.; Li, X.; Wang, H.; PloS one 2014, 9, e110420.
(46) Zhao, N.; Workman, B.; Zhu, D. Endothelialization of novel
Xiao, X.; Li, L.; Zhao, S.; Guo, Z.; Pu, Z.; Zheng, Y. Mechanical
magnesium-rare earth alloys with fluoride and collagen coating. Int. J.
properties, in vitro degradation behavior, hemocompatibility and
Mol. Sci. 2014, 15, 5263−5276.
cytotoxicity evaluation of Zn-1.2Mg alloy for biodegradable implants.
(47) Thevenot, P. T.; Baker, D. W.; Weng, H.; Sun, M.-W.; Tang, L.
RSC Adv. 2016, 6, 86410−86419. The pivotal role of fibrocytes and mast cells in mediating fibrotic
(29) Wang, C.; Yang, H. T.; Li, X.; Zheng, Y. F. In Vitro Evaluation reactions to biomaterials. Biomaterials 2011, 32, 8394−8403.
of the Feasibility of Commercial Zn Alloys as Biodegradable Metals. J. (48) Thevenot, P. T.; Nair, A. M.; Shen, J.; Lotfi, P.; Ko, C.-Y.;
Mater. Sci. Technol. 2016, 32, 909−918. Tang, L. The effect of incorporation of SDF-1α into PLGA scaffolds
(30) Gong, H.; Wang, K.; Strich, R.; Zhou, J. G. In vitro on stem cell recruitment and the inflammatory response. Biomaterials
biodegradation behavior, mechanical properties, and cytotoxicity of 2010, 31, 3997−4008.
biodegradable Zn-Mg alloy. J. Biomed. Mater. Res. B 2015, 103, 1632− (49) Chen, Y.; Zhang, Z.; Yang, K.; Du, J.; Xu, Y.; Liu, S. Myeloid
1640. zinc-finger 1 (MZF-1) suppresses prostate tumor growth through
(31) Vojtech, D.; Kubasek, J.; Serak, J.; Novak, P. Mechanical and enforcing ferroportin-conducted iron egress. Oncogene 2014, 34, 3839.
corrosion properties of newly developed biodegradable Zn-based (50) Pierson, D.; Edick, J.; Tauscher, A.; Pokorney, E.; Bowen, P.;
alloys for bone fixation. Acta Biomater. 2011, 7, 3515−3522. Gelbaugh, J.; Stinson, J.; Getty, H.; Lee, C. H.; Drelich, J.; Goldman,
(32) Zhu, D.; Su, Y.; Young, M. L.; Ma, J.; Zheng, Y.; Tang, L. J. A simplified in vivo approach for evaluating the bioabsorbable
Biological Responses and Mechanisms of Human Bone Marrow behavior of candidate stent materials. J. Biomed. Mater. Res. B 2011,
Mesenchymal Stem Cells to Zn and Mg Biomaterials. ACS Appl. 100, 58−67.
Mater. Interfaces 2017, 9, 27453−27461. (51) Bowen, P. K.; Drelich, J.; Goldman, J. Zinc exhibits ideal
(33) Yang, H.; Wang, C.; Liu, C.; Chen, H.; Wu, Y.; Han, J.; Jia, Z.; physiological corrosion behavior for bioabsorbable stents. Adv. Mater.
Lin, W.; Zhang, D.; Li, W.; Yuan, W.; Guo, H.; Li, H.; Yang, G.; Kong, 2013, 25, 2577−2582.
D.; Zhu, D.; Takashima, K.; Ruan, L.; Nie, J.; Li, X.; Zheng, Y. (52) Bowen, P. K.; Guillory, R. J.; Shearier, E. R.; Seitz, J.-M.;
Evolution of the degradation mechanism of pure zinc stent in the one- Drelich, J.; Bocks, M.; Zhao, F.; Goldman, J. Metallic zinc exhibits
year study of rabbit abdominal aorta model. Biomaterials 2017, 145, optimal biocompatibility for bioabsorbable endovascular stents. Mater.
92−105. Sci. Engin. C 2015, 56, 467−472.
(34) Zhao, S.; McNamara, C. T.; Bowen, P. K.; Verhun, N.; (53) Baker, D. W.; Tsai, Y.-T.; Weng, H.; Tang, L. Alternative
Braykovich, J. P.; Goldman, J.; Drelich, J. W. Structural Characteristics strategies to manipulate fibrocyte involvement in the fibrotic tissue
and In Vitro Biodegradation of a Novel Zn-Li Alloy Prepared by response: pharmacokinetic inhibition and the feasibility of directed-
Induction Melting and Hot Rolling. Metall. Mater. Trans. A 2017, 48, adipogenic differentiation. Acta Biomater. 2014, 10, 3108−3116.
1204−1215. (54) Baker, D. W.; Zhou, J.; Tsai, Y.-T.; Patty, K. M.; Weng, H.;
(35) Bowen, P. K.; Seitz, J. M.; Guillory, R. J., 2nd; Braykovich, J. P.; Tang, E. N.; Nair, A.; Hu, W.-J.; Tang, L. Development of optical
Zhao, S.; Goldman, J.; Drelich, J. W. Evaluation of wrought Zn-Al probes for in vivo imaging of polarized macrophages during foreign
alloys (1, 3, and 5 wt % Al) through mechanical and in vivo testing for body reactions. Acta Biomater. 2014, 10, 2945−2955.
(55) Tsai, Y.-T.; Zhou, J.; Weng, H.; Shen, J.; Tang, L.; Hu, W.-J.
stent applications. J. Biomed. Mater. Res. Part B Appl. Biomater. 2018,
Real-time noninvasive monitoring of in vivo inflammatory responses
106, 245−258.
using a pH ratiometric fluorescence imaging probe. Adv. Healthc.
(36) Feyerabend, F.; Fischer, J.; Holtz, J.; Witte, F.; Willumeit, R.;
Mater. 2013, 3, 221−229.
Drücker, H.; Vogt, C.; Hort, N. Evaluation of short-term effects of
(56) Tsai, Y.-T.; Zhou, J.; Weng, H.; Tang, E. N.; Baker, D. W.;
rare earth and other elements used in magnesium alloys on primary Tang, L. Optical imaging of fibrin deposition to elucidate
cells and cell lines. Acta Biomater. 2010, 6, 1834−1842. participation of mast cells in foreign body responses. Biomaterials
(37) Zeng, R.; Dietzel, W.; Witte, F.; Hort, N.; Blawert, C. Progress 2014, 35, 2089−2096.
and challenge for magnesium alloys as biomaterials. Adv. Eng. Mater. (57) Zhou, J.; Tsai, Y.-T.; Weng, H.; Baker, D. W.; Tang, L. Real
2008, 10, B3−B14. time monitoring of biomaterial-mediated inflammatory responses via
(38) Ma, J.; Zhao, N.; Zhu, D. Sirolimus-eluting dextran and macrophage-targeting NIR nanoprobes. Biomaterials 2011, 32, 9383−
polyglutamic acid hybrid coatings on AZ31 for stent applications. J. 9390.
Biomater. Appl. 2015, 30, 579−588. (58) Yang, H.; Qu, X.; Lin, W.; Wang, C.; Zhu, D.; Dai, K.; Zheng,
(39) Zhao, N.; Watson, N.; Xu, Z.; Chen, Y.; Waterman, J.; Sankar, Y. In vitro and in vivo studies on zinc-hydroxyapatite composites as
J.; Zhu, D. In vitro biocompatibility and endothelialization of novel novel biodegradable metal matrix composite for orthopedic
magnesium-rare Earth alloys for improved stent applications. PloS one applications. Acta Biomater. 2018, 71, 200−214.
2014, 9, e98674. (59) Zhao, S.; Seitz, J.-M.; Eifler, R.; Maier, H. J.; Guillory, R. J.;
(40) Wang, J.; Witte, F.; Xi, T.; Zheng, Y.; Yang, K.; Yang, Y.; Zhao, Earley, E. J.; Drelich, A.; Goldman, J.; Drelich, J. W. Zn-Li alloy after
D.; Meng, J.; Li, Y.; Li, W.; Chan, K.; Qin, L. Recommendation for extrusion and drawing: Structural, mechanical characterization, and
modifying current cytotoxicity testing standards for biodegradable biodegradation in abdominal aorta of rat. Mat. Sci. Eng. C Mater. Biol.
magnesium-based materials. Acta Biomater. 2015, 21, 237−249. Appl. 2017, 76, 301−312.

6818 DOI: 10.1021/acsami.8b20634


ACS Appl. Mater. Interfaces 2019, 11, 6809−6819
ACS Applied Materials & Interfaces Research Article

(60) McCall, K. A.; Huang, C.-c.; Fierke, C. A. Function and


mechanism of zinc metalloenzymes. J. Nutr. 2000, 130, 1437S−
1446S.
(61) Trumbo, P.; Yates, A. A.; Schlicker, S.; Poos, M. Dietary
Reference Intakes. J. Am. Diet. Assoc. 2001, 101, 294−301.
(62) Su, Y.; Cockerill, I.; Wang, Y.; Qin, Y.-X.; Chang, L.; Zheng, Y.;
Zhu, D. Zinc-Based Biomaterials for Regeneration and Therapy.
Trends Biotechnol. 2018, DOI: 10.1016/j.tibtech.2018.10.009, . In
press .
(63) Malekani, J.; Schmutz, B.; Gu, Y.; Schuetz, M.; Yarlagadda, P.
K. In Biomaterials in orthopedic bone plates: a review. Proceedings of
the 2nd Annual International Conference on Materials Science, Metal &
Manufacturing (M3 2011); Global Science and Technology Forum,
2011; pp 71−77.
(64) Mostaed, E.; Sikora-Jasinska, M.; Mostaed, A.; Loffredo, S.;
Demir, A. G.; Previtali, B.; Mantovani, D.; Beanland, R.; Vedani, M.
Novel Zn-based alloys for biodegradable stent applications: Design,
development and in vitro degradation. J. Mech. Behav. Biomed. Mater.
2016, 60, 581−602.
(65) Kubasek, J.; Vojtěch, D. Zn-based alloys as an alternative
biodegradable materials. Proc. Metal 2012, 5, 23−25.

6819 DOI: 10.1021/acsami.8b20634


ACS Appl. Mater. Interfaces 2019, 11, 6809−6819

You might also like