Article Comparative in vitro study and biomechanical testing of two different magnesium alloys Andreas Weizbauer1,2, Christian Modrejewski1, Sabine Behrens3, Helmut Klein4, Patrick Helmecke5, Jan-Marten Seitz3, Henning Windhagen1, Kai Mo¨hwald3, Janin Reifenrath6 and Hazibullah Waizy1 Abstract In this in vitro study, magnesium plates of ZEK100 and MgCa0.8 alloy similar to common titanium alloy osteosynthesis plates were investigated as degradable biomedical materials with a focus on primary stability. Immersion tests were performed in Hank’s Balanced Salt Solution at 37C. The bending strength of the samples was determined using the four- point bending test according to ISO 9585:1990. The initial strength of the noncorroded ZEK100 plate was 11% greater than that of the MgCa0.8 plate; both were approximately 65% weaker than a titanium plate. The bending strength was determined after 48 and 96 h of immersion in Hank’s Balanced Salt Solution; both magnesium alloys decreased by approximately 7% after immersion for 96 h. The degradation rate and the Mg2þ release of ZEK100 were lower than those of MgCa0.8. Strong pitting and filiform corrosion were observed in the MgCa0.8 samples after 96 h of immersion. The surface of the ZEK100 plates exhibited only small areas of filiform corrosion. The results of this in vitro study indicate that the ZEK100 alloy may be more suitable for biomedical applications. Keywords In vitro, corrosion, magnesium, orthopedic, implants Introduction Magnesium and its alloys were rediscovered at the end of the 20th century as promising degradable biomater- ials for orthopedic applications.1–3 Researchers are attracted by the prospect of avoiding a second surgery for implant removal, which is necessary with titanium implants.4 In addition, magnesium alloys provide desir- able mechanical properties, close to those of natural bone, which prevent the ‘‘stress shielding’’ phenomena.5 Magnesium reportedly stimulates new bone formation (osteoinductive) and also supports the adherence of bone tissue to its surface (osteoconductive).5–8 In human body fluids and blood plasma with high Cl– concentrations, the degradation of magnesium alloys is accompanied by a loss of mechanical stability. The resulting surface layer consists mainly of the cor- rosion product magnesium hydroxide (Mg(OH)2), which is locally dissolved into soluble MgCl2. In add- ition, calcium phosphate salts9 and MgCO3 10 precipi- tate on the surface of magnesium alloys resulting in a corrosive protective effect. During the corrosion pro- cess, Mg2þ ions are released from the metal. Magnesium is a natural component of the human body and it is mainly stored in bone (60–65%).11 Furthermore, it is the second most abundant cation in cells and an important cofactor for many enzymes that are involved in metabolic pathways and DNA repair processes.11,12 The concentration of Mg2þ in the blood serum of human adults comprises 0.7–1.1mol/l; Journal of Biomaterials Applications 2014, Vol. 28(8) 1264–1273 ! The Author(s) 2013 Reprints and permissions: sagepub.co.uk/journalsPermissions.nav DOI: 10.1177/0885328213506758 jba.sagepub.com 1Department of Orthopedic Surgery, Hannover Medical School, Hannover, Germany 2CrossBIT, Center for Biocompatibility and Implant-Immunology, Department of Orthopedic Surgery, Hannover Medical School, Hannover, Germany 3Institute of Materials Science, Leibniz Universita¨t Hannover, Garbsen, Germany 4Department of Crystallography, University of Go¨ttingen, Go¨ttingen, Germany 5Institute of Production Engineering and Machine Tools (IFW), Leibniz Universita¨t Hannover, Garbsen, Germany 6Small Animal Clinic, University of Veterinary Medicine Hannover, Hannover, Germany Corresponding author: Andreas Weizbauer, Department of Orthopedic Surgery, Hannover Medical School, Anna-von-Borries-Str. 1-7, Hannover 30625, Germany. Email: weizbauer.andreas@mh-hannover.de at SUB Goettingen on September 3, 2014jba.sagepub.comDownloaded from serum Mg2þ levels exceeding this amount are excreted by the kidneys.11 In this in vitro study, two different magnesium alloys were investigated: one containing rare earth elements (REEs) (ZEK100) and MgCa0.8. Both alloys were chosen because of their promising biomechanical prop- erties, and both had been evaluated in vivo in previous studies.13–21 ZEK100 pins have shown promising initial mechanical stability.15,20 Lensing et al. reported good biocompatibility and an osteoconductive effect of ZEK100 in the middle ear of rabbit.17 In contrast, some studies have shown that the biocompatibility of the ZEK100 alloy appears to be questionable.16,22 Pure magnesium possesses insufficient mechanical properties.23 The alloying of magnesium with a large variety of elements was introduced to solve this issue. These alloying elements influence the corrosion film composition and therefore the corrosion properties.24 Magnesium alloyed with passivating elements, such as the REE (yttrium and neodymium), exhibits enhanced corrosion resistance.18,24 Magnesium alloys containing REE were first investigated at the beginning of the 1980 s.25 The addition of REE contributes to solid solu- tion strengthening and grain boundary strengthening, and therefore enhances the ductility of the alloy.1,25 Several studies have investigated the properties and corrosion behavior of binary magnesium–calcium alloys.13,14,18,19,26–30 The addition of calcium provides solid solution strengthening and contributes to grain refinement owing to grain boundary strengthening.1 It was observed that the addition of calcium above the solubility limit of approximately 1.34wt% Ca enhanced the corrosion rate.18,19,26 The addition of less than 1wt% Ca demonstrated a low corrosion rate with an appropriate elastic modulus.18 Previous studies have demonstrated good biocom- patibility of MgCa0.8 in vivo.13,28 The production of high amounts of gas in a short time due to high corro- sion rates is undesirable for clinical applications. The accumulation of gas associated with MgCa0.8 implants was mild to moderate in vivo, with no harmful effects toward the animal.14 In this in vitro study, we investigated the corrosion behavior and mechanical properties of two different alloys (MgCa0.8 and ZEK100) in Hank’s Balanced Salt Solution (HBSS). The results were compared with a titanium alloy plate (Ti6Al4V) of similar sample size and shape. This study is focused on primary stability and the effects of degradation during the first days of immersion. In addition, Mg2þ release from the alloy was determined, and the corrosion rate was eval- uated according to the mass loss method. Alloy sur- faces were investigated using scanning electron microscopy (SEM), and high-energy synchrotron radi- ation was applied for microstructural characterization. Methods Material and preparation Two different magnesium alloys were examined in this study: ZEK100 (1wt% zinc, 0.1wt% zirconium, and 0.1wt% rare earth metals) and MgCa0.8 (0.8wt% cal- cium). The plates were 50mm long, 8mm wide, and 2mm thick. They were manufactured by face slab milling. Five screw holes were drilled and counterbored in each plate. The resulting screw holes had an inner diameter of 3mm on the bone-facing side and a counterbore diameter of 5mm (Figure 1). The samples had a surface of 898mm2. Two titanium alloy plates with the same dimensions were manufactured from a Ti6Al4V rod (Goodfellow GmbH, Bad Nauheim, Germany). The surfaces of the magnesium samples were pickled prior to immersion testing with a solution of 5ml 65% nitric acid (Merck, Darmstadt, Germany), 20ml 85% glycerol (Merck, Darmstadt, Germany), and 5ml 100% acetic acid (Merck, Darmstadt, Germany) for 30 s. Next, the samples were ultrasonically cleaned for 2min with isopropanol and then dried in a vacuum drying oven (Heraeus Vacuum oven, Thermo Scientific, Bonn, Germany) at 50C for approximately 20min. Immersion testing The static immersion test was performed in HBSS (Biochrom AG, Berlin, Germany) to simulate the normal ion concentration under physiological condi- tions at 37.5C. The corrosion media contained 8.00 g/l NaCl, 0.4 g/l KCl, 0.04 g/l Na2HPO4, 0.06 g/l KH2PO4, 0.20 g/l MgSO47H2O, 0.14 g/l CaCl2, 1.00 g/l glucose, and 0.35 g/l NaHCO3. The plates were placed into 500ml HBSS by a nonresorbable Ti6Al4V MgCa0 .8 ZEK10 0 Figure 1. Picture of the plates (length: 50mm, width: 8mm, thickness: 2mm) used in this in vitro study. Weizbauer et al. 1265 at SUB Goettingen on September 3, 2014jba.sagepub.comDownloaded from wire (EthiconTMProleneTM, 4-0, Johnson & Johnson GmbH, Neuss, Germany). The pH of the solution was monitored during the immersion test using a pH meter (Schott Instruments Lab850, SI Analytics GmbH, Mainz, Germany). After immersion, the sam- ples were dried and the corrosion layer was removed with chromic acid. Two samples were used per interval of each magnesium alloy.The degradation rate was evaluated according to the weight loss method31 W ¼Wbefore Wafter At where W is the weight loss, Wbefore (mg) is the weight of each sample before corrosion, and Wafter (mg) is the weight of each sample after corrosion. A is defined as the sample’s surface area (cm2) and t is the immersion time (d). This weight loss rate was converted into an average corrosion rate (mm/y) using31,32 PW ¼ 2:1W Detection of Mg2þ release through inductively coupled plasma optical emission spectroscopy (ICP-OES) The ICP-OES (Spectro Ciros Vision EOP, Spectro Analytical Instruments, Kleve, Germany) was applied to quantitatively determine the release of magnesium ions from the plates during the corrosion process. Two plates of each alloy were immersed in 100ml HBSS at 37C for 158 h. At fixed time intervals (the first after 45min), 1ml was taken for analysis and replaced with the same amount of fresh corrosion media. The pH of every analysis sample was monitored. Biomechanical testing To specify the mechanical properties of the specimens, they were tested according to ISO 9585:1990. The four- point bending test setup (Figure 2) was designed for tests with a uniaxial material testing machine (Mini Bionix 858, MTS Systems in Minneapolis, MN, USA). The test system consisted of four rollers with a diameter of 8mm. The distance between the lower and outer rollers was 32mm, and the distance between the upper and inner rollers was 10mm. The plates were positioned on the lower rollers according to the proto- col of ISO 9585:1990. The distance between the outer and inner rollers was 11mm, including a screw hole of the demonstrators. The central screw hole and another screw hole were positioned between the inner rollers. The test rig was produced by the Research Workshop of the Medical School, Hannover, Germany. Microstructural characterization After the corrosion products were removed from the surface with chromic acid, the magnesium plates were analyzed with a Zeiss EVO 60 VP SEM. Diffraction experiments were performed using high- energy synchrotron radiation (E100 keV, ¼ 0.12561 A˚). A Perkin–Elmer area detector (400mm 400mm, 200 mm pixel size) was used with a sample–detector distance of 1430mm. Results Immersion test and corrosion properties The pH of HBSS was measured during the immersion of the plates (Figure 3). The pH of the HBSS increased rapidly during the first hours of exposure. After 12 h of Figure 2. Picture of the four-point bending setup. The samples were tested according to ISO 9585:1990. 1266 Journal of Biomaterials Applications 28(8) at SUB Goettingen on September 3, 2014jba.sagepub.comDownloaded from immersion, the alkalization of the corrosion medium progressed at a slower pace. However, the immersion of MgCa0.8 plates resulted in a higher HBSS pH than the immersion of ZEK100 plates. After 96 h of immer- sion, the pH of HBSS rose to 9.9 for MgCa0.8 plates compared with 8.8 for ZEK100 plates. The corrosion rate was determined after 48 and 96 h using the weight loss method (Table 1). ZEK100 demonstrated a five-fold lower corrosion rate than MgCa0.8. Release of Mg2þ from the magnesium plates The Mg2þ concentration dissolved into the HBSS from the magnesium samples was determined using ICP-OES (Figure 4(a)). The MgCa0.8 plates exhibited rapid Mg2þ release during the first 20 h of immersion. The ZEK100 plates exhibited a more consistent increase of Mg2þ concentration throughout the exposure time. The rate of Mg2þ release from the MgCa0.8 plates during the first 24 h of immersion was 1.082mg/h (1.2 10–3mg/hmm2), nearly eight-fold higher than the rate of ion release from the ZEK100 plates: 0.143mg/h (1.592 10–4mg/hmm2) (Figure 4(b)). Moreover, the pH of HBSS increased more rapidly as a result of the immersion of the MgCa0.8 samples than the ZEK100 samples. After 20 h, the MgCa0.8 corrosion media reached a stable pH of approximately 10 (Figure 4(c)). The effect of corrosion on bending strength Figure 5 shows the bending strengths of the magnesium and titanium plates. Uncorroded samples of ZEK100 (0.306 0.003Nm) were stronger than MgCa0.8 (0.272 0.00005Nm). In contrast, the bending strength of titanium plates with the same dimensions was higher (0.803 0.001N m). The strength of both magnesium alloys decreased with immersion time. The bending strength after 96 h of immersion was approximately 7% lower than the initial bending strength (0 h immersion). Microstructural characterization and phase identification Figure 6 shows the surface of the magnesium plates after removal of the surface corrosion products that resulted from 96 h of immersion. Filiform and pitting corrosion were observed on the surface of the MgCa0.8 alloy. The surface of the ZEK100 plates showed only small areas of filiform corrosion (Figure 6(a) and (b)). The diffraction diagram and the diffraction rings of both alloys are displayed in Figure 7. The diffraction rings of both alloys demonstrated an inhomogeneous intensity distribution (Figure 7(b) and (e)). The diffrac- tion diagram of the MgCa0.8 alloy is shown in Figure 7(c) and (d). In addition to the measured intensities 10.50 10.00 9.50 9.00 8.50 8.00 7.50 0 20 40 60 Immersion time (h) pH v al ue 80 100 MgCa0.8 ZEK100 Figure 3. pH value versus immersion time of MgCa0.8 and ZEK100 plates in 500ml Hank’s solution. The dots represent mean values and standard deviation. Sample size for each alloy: n¼ 4 for 0–48 h, n¼ 2 each for 72 and 96 h. Table 1. Corrosion rates of the magnesium alloy plates according to the mass loss method.31 Sample size: Two of each alloy per interval. Mean corrosion rate (mm/y) 48 h immersion 96 h immersion MgCa0.8 3.62 2.89 ZEK100 0.70 0.63 Weizbauer et al. 1267 at SUB Goettingen on September 3, 2014jba.sagepub.comDownloaded from (black line), simulated diffraction diagrams are also provided for Mg (solid solution, red line) and Mg2Ca (intermetallic compound, blue line). Both simulated phases fit the position of the measured reflection lines quite well, but not the intensity. The measured intensi- ties of the ZEK100 samples also differed strongly because of the material’s texture. No intermetallic phases were detected. 60(a) (b) (c) 50 40 30 20 10 0 0 10 20 30 40 50 60 70 80 90 100 110 120 130 140 150 MgCa0.8 ZEK100 MgCa0.8 ZEK100 Linear regression line MgCa.8 y =0.143x + 15.029 y =1.082x + 10.846 Linear regression line ZEK100 MgCa0.8 ZEK100 0 10 20 30 40 50 60 0 4 8 12 16 20 24 70 80 Immersion time (h) Immersion time (h) Immersion time (h) 90 100110120130140150 45 40 35 30 25 20 15 5 10 0 10.5 pH v al ue M g2 + -c on ce nt ra tio n (m g/ l) M g2 + -c on ce nt ra tio n (m g/ l) 10 9.5 9 8.5 8 7.5 Figure 4. (a) Mg2þ release from the ZEK100 and MgCa0.8 plates as a function of time. (b) Average Mg2þ release during the first 24 h. (c) pH of the analyzed corrosion media sample. The dots represent mean values and standard deviation. Two samples were tested per alloy. 0.805 titanium plates without immersion MgCa0.8 ZEK100 0.305 0.295 0.275 0.285 Be nd in g st re ng th (N m) 0.265 0.255 0.245 0 48 Immersion time (h) 96 Figure 5. Bending strength versus immersion time. The bending strength of both alloys decreased with immersion time. The dots represent mean values and standard deviation. Two samples were tested per alloy and immersion time. 1268 Journal of Biomaterials Applications 28(8) at SUB Goettingen on September 3, 2014jba.sagepub.comDownloaded from 0 200 400 600 800 1000 1200 1400 1600 1800 2000 0 200 400 600 800 1000 Columns Columns 1200 1400 1600 1800 2000 0 200 400 600 800 1000 1200 1400 1600 1800 2000 0 0 200 200 400 400 600 600 800 800 1000 1000 1200 1200 1400 1400 1600 1600 1800 1800 2000 2000 0 200 400 600 800 1000 1200 1400 1600 1800 2000 0 200 400 600 800 1000 1200 1400 1600 1800 2000 0 200 400 600 800 1000 R ow s R ow s 1200 1400 1600 Mg Measurement Mg Measurement Mg 2 Ca 1800 2000 1 2 3 4 5 6 7 8 9 10 1 2 3 4 5 2 Theta/° 2 Theta/°3 0 4 2 Theta/° In te ns ity In te ns ity In te ns ity 6 7 8 9 10 (a) (b) (d)(c) (d) (d) Figure 7. Diffraction diagram and diffraction rings of ZEK100 and Mg2Ca0.8 alloy samples. (a) and (b): ZEK100 and (c) to (e): Mg2Ca0.8. Figure 6. SEM pictures of the plates after 96 h immersion and removal of the corrosion layer. (a) and (b): ZEK100 and (c) and (d): MgCa0.8. Weizbauer et al. 1269 at SUB Goettingen on September 3, 2014jba.sagepub.comDownloaded from Discussion Fracture healing is dependent on the fracture gap and achieved stabilization. Plates are a common orthopedic implant and are usually attached to the bone with screws.33 The fixation of a plate should supply sufficient stability to enhance fracture healing. High primary sta- bility is important and should not decrease greatly during the first days after the placement of the plate. Therefore, in this in vitro study we focused on the ini- tial strength of the magnesium implants versus titanium implants and investigated the implant stability specific- ally during the first days of immersion. The currently used plates are made of stainless steel or titanium alloys5 and are available in a wide range of sizes. In this study, five-hole plates from two different mag- nesium alloys were examined after immersion in HBSS and compared with titanium plates. The biomechanical results of the magnesium-based plates were compared with those of a plate made from the titanium alloy Ti6Al4V, which is commonly used in orthopedic trauma surgery. Titanium alloys exhibit minor biocorrosion with the release of metallic ions, which normally do not compromise plate stability.34 The corrosion products, released metallic ions, have been associated with proinflammatory, cytokine- induced, enhanced osteoclastic activity, which in turn causes bone resorption.34 We assume that the loss of stability of titanium alloys is negligible within the short period of time (96 h) observed in our study. Therefore, the biomech- anical examinations of the titanium plates were per- formed to their initial conditions without immersion tests. The bending strength of the titanium plates is com- parable to that of a 2mm mini-fragment plate, which was also tested according to ISO 9585:1990.35 In this study a bending strength of 0.53Nm was determined. The biomechanical parameter of the tested 2mm mini- fragment plate is lower than the bending strength of the titanium plates evaluated in our study (0.803Nm). The reason for this slight deviation might be explained by the difference in plate design. The initial strengths of both magnesium alloys were approximately 65% lower than that of the titanium alloy. Of the two magnesium alloys, the ZEK100 plates exhibited greater initial stability. Therefore, ZEK100 plates may be considered more suitable for osteosynthesis applications than MgCa0.8 plates. The degradation of magnesium alloys is known to be high in the first hours of exposure to the corrosion media. This is mainly due to the fact that the corrosion protective layer needs some time for formation. Therefore, we particularly investigated the first hours of corrosion and its effect on the stability of the magnesium plates. The study of Waizy et al. showed that between week 2 and week 6 of immersion of ZEK100 in HBSS no con- siderable change in bending strength was reported.36 Our study showed that in the first 96 h of degradation the bending strength of both magnesium alloys decreased about 7%. ZEK100 still have a higher stabil- ity and is therefore based on the biomechanical results of the preferred magnesium alloy. Many studies have shown that the degradation of magnesium alloys in in vitro experiments is higher than in comparable in vivo studies.37–39 A decreased loss of stability in vivo com- pared to the results in this study is assumed. Alloy plates are used in clinical practice as internal splints to hold fractured ends of bone together; add- itional fracture stabilization with lag screws is common. According to the biomechanical results, the preferred use of these magnesium-based plates is the fixation of non-weight-bearing bones (e.g. fracture of the metacarpal bones). Sufficient stability of degradable plates in vivo must be ensured during the first 6–8 weeks after implantation.40 The magnesium alloy ZEK100 contains REE and zirconium in low concentrations. The alloying elements are used to improve the mechanical and corrosion properties of the biomaterial.41 These alloying elements are known to have adverse effects and potential long- term risks,42–45 although some studies have reported good biocompatibility.46,47 Allergic hypersensitivity, reduction of hemoglobin oxygen affinity, and fatty liver are some potential adverse effects of these elements at toxic doses.43,45 Chelated REE are easily excreted in the urine, and mortality studies have revealed that REE are not highly toxic.43 However, bone is one of the main target organs of REE because REE are of a radius similar to Ca2þ ions, which may lead to REE accumulation in the environment directly surrounding the implant.43 Because implants made of degradable magnesium alloys are not removed after implantation, in vivo biocompatibility studies should focus on the long-term effects of these implants. In an animal study with 12 months of observation time, a good bio- logical response to a REE-containing alloy (MgYREZr, similar to WE43) and its degradation products was observed.48 The degradation of magnesium and its alloys in bio- fluids is described by the following reaction Anodic reaction: Mg ! Mg2þþ 2 e- Cathodic reaction: 2H2Oþ 2 e– ! H2"þ 2OH– Mg2þ þ 2OH !Mg OHð Þ2 sð Þ Degradation is accompanied by an alkalization of the corrosion media due to the production of hydroxide ions (OH–). These experiments were performed without CO2 gas application. HCO3 –/CO2 is the main buffering 1270 Journal of Biomaterials Applications 28(8) at SUB Goettingen on September 3, 2014jba.sagepub.comDownloaded from system in blood and in HBSS. The high proportion of hydroxide ions supports the formation of magnesium hydroxide, which in turn acts as a protective layer against corrosion. The greater tendency toward film formation is reportedly associated with local pH> 10.5 and is believed to have a corrosion-inhibiting effect.49 Magnesium hydroxide is disrupted by chloride ions with the release of OH– Mg OHð Þ2þ2Cl !MgCl2 þ 2OH In response to the immersion of both magnesium alloys (ZEK100 and MgCa0.8), the pH of the HBSS increased sharply during the first hours of exposure. An acceler- ated pH increase during the first hours of immersion was reported by several in vitro studies.50–53 The decel- erated increase of pH after 6 h may be attributed to the formation of magnesium hydroxide and other phos- phate- and calcium-containing salts, which precipitated on the surface of the magnesium samples. Interestingly, the pH of the HBSS increased more with the immersion of MgCa0.8 plates than ZEK100 plates. This finding indicated the more progressed and faster corrosion of the MgCa0.8 plates. These observations were sup- ported by the determined corrosion rates for both mag- nesium alloys according to the mass loss method (Table 1). The corrosion rates were approximately five-fold greater for the MgCa0.8 plates than for the ZEK100 plates. The corrosion rate was slower after 96 h of immersion than after 48 h, which can be explained by the protective effect of the formed surface films (result- ing from the precipitation of calcium phosphate salts and magnesium hydroxide). Song et al. reported the formation of a magnesium hydroxide film on the sur- face of an AZ31 within 2 h after immersion.23 Similar to our results, Zhu et al. reported a reduction of the cor- rosion rate with progressing immersion time for a AZ31 alloy.54 A study by Walker et al. reported lower corro- sion rates for MgCa0.8 than were observed in the pre- sent study (1.937mm/y in MEMp, 1.291mm/y in MEM, and 0.795mm/y in EBSS).38 Several studies have shown that the corrosion rates in vivo differ from those determined in vitro.39,55 Many different parameters influence the in vitro corrosion rate (e.g. choice of corrosion media, temperature, CO2 gas application, etc.).3,56,57 Therefore, the alteration of parameters makes it difficult to compare different in vitro tests. The release of Mg2þ from the metal was measured using ICP-OES analysis (Figure 4). Only a few studies have analyzed the release of Mg2þ from magnesium alloys.51,52 These results confirm the observations of the immersion test and the determination of the corro- sion rate. The release of Mg2þ decelerates with immer- sion time as a consequence of protective surface layer formation. This analysis is influenced by the formation and degradation of Mg(OH)2. Some of the dissolved Mg2þ ions from the metal form insoluble magnesium hydroxide Mg(OH)2, and therefore are excluded from the analysis, although as mentioned previously, Mg(OH)2 can be transformed into soluble MgCl2 by Cl– ions and is therefore again accessible for ICP- OES analysis. However, the speed of this accumulation and degradation process is unknown. Notably, the high rate of MgCa0.8 corrosion appears to result in a faster loss of biomechanical stability. The microstructural characterization of the MgCa0.8 plates revealed the presence of the intermetal- lic Mg2Ca phase, which is consistent with other studies.26,58 Intermetallic phases form during the solidi- fication of the cast alloys and mainly distribute at the grain boundaries.59 Most of the intermetallic second phases reportedly possess a corrosion-accelerating effect.60 This reduction of corrosion resistance is mainly attributed to galvanic corrosion between the anodic a-Mg and the intermetallic phases, which act as a cathode.60 Kirkland et al. report that this phase is very reactive, with faster dissolution rates than Mg.26 The Mg2Ca phase is described as a more efficient anode than a-Mg.26 Analysis of the ZEK100 plate revealed no interme- tallic phases. Here, a combined Rietveld texture ana- lysis would have given very precisely the amounts of the present phases and the texture (to calculate the aniso- tropic behavior of the material). Small amounts of other phases might also be found using this method, because of the improved alignment of the measured intensities. MgZn and Mg2Zn are the most prominent intermetallic phases that would be expected according to an analysis of a similar magnesium alloy.51,61 Conclusion Magnesium plates made of different magnesium alloys were investigated after in vitro corrosion in HBSS. Biomechanical tests demonstrated a continuous loss of mechanical strength by osteosynthesis plates made of magnesium alloys. Although the initial bending strength of the uncorroded ZEK100 plate was greater than that of the MgCa0.8 plate, both were approxi- mately 65% weaker than a titanium plate. The degrad- ation rate and the Mg2þ release of ZEK100 were lower than those of MgCa0.8. Acknowledgments The authors gratefully acknowledge the financial support pro- vided by the German Research Society (DFG) within the col- laborative research project (SFB 599). We thank Markus Badenhop for excellent technical support and Christopher Mu¨ller for the design of Figure 1. Weizbauer et al. 1271 at SUB Goettingen on September 3, 2014jba.sagepub.comDownloaded from References 1. Witte F, Hort N, Vogt C, et al. Degradable biomaterials based on magnesium corrosion. Curr Opin Solid State Mater Sci 2008; 12(5–6): 6372. 2. Witte F. The history of biodegradable magnesium implants: A review. Acta Biomater 2010; 6(5): 16801692. 3. Waizy H, Seitz J-M, Reifenrath J, et al. Biodegradable magnesium implants for orthopedic applications. J Mater Sci 2013; 48(1): 3950. 4. Yang L, Huang Y, Peng Q, et al. Mechanical and corro- sion properties of binary Mg-Dy alloys for medical appli- cations. Mater Sci Eng B 2011; 176(20): 18271834. 5. Staiger MP, Pietak AM, Huadmai J, et al. Magnesium and its alloys as orthopedic biomaterials: A review. Biomaterials 2006; 27(9): 17281734. 6. Albrektsson T and Johansson C. Osteoinduction, osteo- conduction and osseointegration. Eur Spine J 2001; 10(Suppl 2): 96101. 7. Xu L, Yu G, Zhang E, et al. In vivo corrosion behavior of Mg-Mn-Zn alloy for bone implant application. J Biomed Mater Res A 2007; 83(3): 703711. 8. Zreiqat H, Howlett CR, Zannettino A, et al. Mechanisms of magnesium-stimulated adhesion of osteoblastic cells to commonly used orthopaedic implants. J Biomed Mater Res 2002; 62(2): 175184. 9. Rettig R and Virtanen S. Composition of corrosion layers on a magnesium rare-earth alloy in simulated body fluids. J Biomed Mater Res A 2009; 88(2): 359369. 10. Willumeit R, Fischer J, Feyerabend F, et al. Chemical surface alteration of biodegradable magnesium exposed to corrosion media. Acta Biomater 2011; 7(6): 27042715. 11. Vormann J. Magnesium: Nutrition and metabolism. Mol Aspects Med 2003; 24(1–3): 2737. 12. Hartwig A. Role of magnesium in genomic stability. Mutat Res 2001; 475(1–2): 113121. 13. Erdmann N, Bondarenko A, Hewicker-Trautwein M, et al. Evaluation of the soft tissue biocompatibility of MgCa0.8 and surgical steel 316L in vivo: A comparative study in rabbits. Biomed Eng Online 2010; 9: 63. 14. Erdmann N, Angrisani N, Reifenrath J, et al. Biomechanical testing and degradation analysis of MgCa0.8 alloy screws: A comparative in vivo study in rabbits. Acta Biomater 2011; 7(3): 14211428. 15. Huehnerschulte TA, Angrisani N, Rittershaus D, et al. In vivo corrosion of two novel magnesium alloys ZEK100 and AX30 and their mechanical suitability as biodegradable implants. Materials 2011; 4(6): 11441167. 16. Huehnerschulte TA, Reifenrath J, von Rechenberg B, et al. In vivo assessment of the host reactions of the bio- degradation of the two novel magnesium alloys ZEK100 and AX30 in an animal model. Biomed Eng Online 2012; 11: 14. 17. Lensing R, Behrens P, Mu¨ller PP, et al. In vivo testing of a bioabsorbable magnesium alloy serving as total ossicu- lar replacement prostheses. J Biomater Appl. Epub ahead of print 2013. DOI: 10.1177/0885328212472246. 18. Li Y, Hodgson PD and Wen C. The effects of calcium and yttrium additions on the microstructure, mechanical properties and biocompatibility of biodegradable magne- sium alloys. J Mater Sci 2011; 46(2): 365371. 19. Li Z, Gu X, Lou S, et al. The development of binary Mg- Ca alloys for use as biodegradable materials within bone. Biomaterials 2008; 29(10): 13291344. 20. Reifenrath J, Bormann D and Meyer-Lindenberg A. Magnesium alloys as promising degradable implant materials in orthopaedic research. In: Czerwinski F (ed.)Magnesium alloys—corrosion and surface treatments. Rijeka: Intech, 2011, pp.93–108. 21. von der Ho¨h N, Bormann D, Lucas A, et al. Influence of different surface machining treatments of resorbable magnesium alloy implants on degradation—EDX-analy- sis and histology results.Materwiss Werkst 2009; 40(1–2): 8893. 22. Dziuba D, Meyer-Lindenberg A, Seitz JM, et al. Long- term in vivo degradation behaviour and biocompatibility of the magnesium alloy ZEK100 for use as a biodegrad- able bone implant. Acta Biomater. Epub ahead of print 2012. DOI: 10.1016/j.actbio.2012.08.028. 23. Song Y, Shan D, Chen R, et al. Biodegradable behaviors of AZ31 magnesium alloy in simulated body fluid. Mater Sci Eng C 2009; 29(3): 10391045. 24. Song G and Atrens A. Understanding magnesium corro- sion—a framework for improved alloy performance. Adv Eng Mater 2003; 5(12): 837858. 25. Al-Samman T and Li X. Sheet texture modification in magnesium-based alloys by selective rare earth alloying. Mater Sci Eng A 2011; 528(10–11): 38093822. 26. Kirkland NT, Birbilis N, Walker J, et al. In-vitro dissol- ution of magnesium-calcium binary alloys: Clarifying the unique role of calcium additions in bioresorbable magne- sium implant alloys. J Biomed Mater Res B Appl Biomater 2010; 95(1): 91100. 27. Krause A, von der Ho¨h N, Bormann N, et al. Degradation behaviour and mechanical properties of magnesium implants in rabbit tibiae. J Mater Sci 2010; 45(3): 624632. 28. Thomann M, Krause C, Bormann D, et al. Comparison of the resorbable magnesium alloys LAE442 and MgCa0.8 concerning their mechanical properties, their progress of degradation and the bone-implant-contact after 12 months implantation duration in a rabbit model. Materwiss Werkst 2009; 40(1–2): 8287. 29. von der Ho¨h N, Bormann D, Lucas A, et al. Influence of different surface machining treatments of magnesium- based resorbable implants on the degradation behavior in rabbits. Adv Eng Mater 2009; 11(5): B47B54. 30. Wan Y, Xiong G, Luo H, et al. Preparation and charac- terization of a new biomedical magnesium-calcium alloy. Mater Design 2008; 29(10): 2034–2037. 31. Zainal Abidin NI, Atrens AD, Martin D, et al. Corrosion of high purity Mg, Mg2Zn0.2Mn, ZE41 and AZ91 in Hank’s solution. Corros Sci 2011; 53(3): 862872. 32. Shi Z, Liu M and Atrens A. Measurement of the corro- sion rate of magnesium alloys using Tafel extrapolation. Corros Sci 2010; 52(2): 579588. 1272 Journal of Biomaterials Applications 28(8) at SUB Goettingen on September 3, 2014jba.sagepub.comDownloaded from 33. Bartel DL, Davy DT and Keaveny TM. Orthopaedic bio- mechanics: Mechanics and design in musculosceletal sys- tems. Englewood Cliffs, NJ: Prentice Hall, 2006. 34. Cadosch D, Chan E, Gautschi OP, et al. Metal is not inert: Role of metal ions released by biocorrosion in asep- tic loosening—current concepts. J Biomed Mater Res A 2009; 91(4): 12521262. 35. Cesarone DM and Disegi JA. Techniques in the applica- tion of ISO 9585 test method for the determination of bone plate bending properties. In: Harvey JP and Games RF (eds) Clinical and laboratory performance of bone plates. West Conshohocken, PA: ASTM International, 1994, pp.65–71. 36. Waizy H, Weizbauer A, Modrejewski C, et al. In vitro corrosion of ZEK100 plates in Hank’s balanced salt solu- tion. Biomed Eng Online 2012; 11: 12. 37. Mueller W-D, Nascimento ML and Lorenzo de Mele MF. Critical discussion of the results from different cor- rosion studies of Mg and Mg alloys for biomaterial appli- cations. Acta Biomater 2010; 6(5): 17491755. 38. Walker J, Shadanbaz S, Kirkland NT, et al. Magnesium alloys: Predicting in vivo corrosion with in vitro immer- sion testing. J Biomed Mater Res B Appl Biomater 2012; 100(4): 11341141. 39. Witte F, Fischer J, Nellesen J, et al. In vitro and in vivo corrosion measurements of magnesium alloys. Biomaterials 2006; 27(7): 10131018. 40. Baumgaertel F, Buhl M and Rahn BA. Fracture healing in biological plate osteosynthesis. Injury 1998; 29(Suppl 3): C3C6. 41. Gu XN and Zheng YF. A review on magnesium alloys as biodegradable materials. Front Mater Sci China 2010; 4(2): 111–115. 42. Haley TJ. Pharmacology and toxicology of the rare earth elements. J Pharm Sci 1965; 54(5): 663670. 43. Hirano S and Suzuki KT. Exposure, metabolism, and toxicity of rare earths and related compounds. Environ Health Perspect 1996; 104(Suppl 1): 8595. 44. Yumiko N, Yukari T, Yasuhide T, et al. Differences in behavior among the chlorides of seven rare earth elem- ents administered intravenously to rats. Fundam Appl Toxicol 1997; 37(2): 106116. 45. Yuen CK and Ip WY. Theoretical risk assessment of magnesium alloys as degradable biomedical implants. Acta Biomater 2010; 6(5): 18081812. 46. Drynda A, Deinet N, Braun N, et al. Rare earth metals used in biodegradable magnesium-based stents do not interfere with proliferation of smooth muscle cells but do induce the upregulation of inflammatory genes. J Biomed Mater Res A 2009; 91(2): 360369. 47. Feyerabend F, Fischer J, Holtz J, et al. Evaluation of short-term effects of rare earth and other elements used in magnesium alloys on primary cells and cell lines. Acta Biomater 2010; 6(5): 18341842. 48. Waizy H, Diekmann J, Weizbauer A, et al. In vivo study of a biodegradable orthopedic screw (MgYREZr-alloy) in a rabbit model for up to 12 months. J Biomater Appl. Epub ahead of print 2013. DOI: 10.1177/ 0885328212472215. 49. Song GL and Atrens A. Corrosion mechanisms of mag- nesium alloys. Adv Eng Mater 1999; 1(1): 1133. 50. Gu XN, Li N, Zhou WR, et al. Corrosion resistance and surface biocompatibility of microarc oxidation coat- ing on a Mg-Ca alloy. Acta Biomater 2011; 7(4): 18801889. 51. Gu XN, Li N, Zheng YF, et al. In vitro degradation performance and biological response of a Mg-Zn-Zr alloy. Mater Sci Eng B 2011; 176(20): 17781784. 52. Huan ZG, Leeflang MA, Zhou J, et al. In vitro degrad- ation behavior and cytocompatibility of Mg-Zn-Zr alloys. J Mater Sci Mater Med 2010; 21(9): 26232635. 53. Yang L and Zhang E. Biocorrosion behavior of magne- sium alloy in different simulated fluids for biomedical application. Mater Sci Eng C 2009; 29(5): 16911696. 54. Zhu Y, Wu G, Zhang Y-H, et al. Growth and character- ization of Mg(OH)2 film on magnesium alloy AZ31. Appl Surf Sci 2011; 257(14): 61296137. 55. Bobe K, Willbold E, Morgenthal I, et al. In vitro and in vivo evaluation of biodegradable, open-porous scaf- folds made of sintered magnesium W4 short fibres. Acta Biomater. Epub ahead of print 2013. DOI: 10.1016/j.actbio.2013.03.035. 56. Willumeit R, Feyerabend F and Huber N. Magnesium degradation as determined by artificial neural networks. Acta Biomater. Epub ahead of print 2013. DOI: 10.1016/ j.actbio.2013.02.042. 57. Xin Y, Hu T and Chu PK. In vitro studies of biomedical magnesium alloys in a simulated physiological environ- ment: A review. Acta Biomater 2011; 7(4): 14521459. 58. Harandi SE, Idris MH and Jafari H. Effect of forging process on microstructure, mechanical and corrosion properties of biodegradable Mg-1Ca alloy. Mater Design 2011; 32(5): 25962603. 59. Hort N, Huang Y, Fechner D, et al. Magnesium alloys as implant materials—principles of property design for Mg- RE alloys. Acta Biomater 2010; 6(5): 17141725. 60. Atrens A, Liu M and Zainal Abidin NI. Corrosion mech- anism applicable to biodegradable magnesium implants. Mater Sci Eng B 2011; 176(20): 16091636. 61. Zhang S, Zhang X, Zhao C, et al. Research on an Mg-Zn alloy as a degradable biomaterial. Acta Biomater 2010; 6(2): 626640. Weizbauer et al. 1273 at SUB Goettingen on September 3, 2014jba.sagepub.comDownloaded from