Вы находитесь на странице: 1из 8

journal of the mechanical behavior of biomedical materials 21 (2013) 1 –8

Available online at www.sciencedirect.com

www.elsevier.com/locate/jmbbm

Research paper

Synthesis, microstructure and mechanical properties


of porous Mg–Zn scaffolds

Z.S. Seyedraoufin, Sh. Mirdamadi


School of Metallurgy and Materials Engineering, Iran University of Science and Technology, 16846-13114 Tehran, Iran

art i cle i nfo ab st rac t

Article history: Magnesium alloys have been intensively studied as biodegradable implant materials, as
Received 16 June 2012 their mechanical properties render them promising candidates for bone tissue engineering
Received in revised form applications. In the present work, porous Mg–4 wt% Zn and Mg–6 wt% Zn scaffolds were
19 January 2013 prepared using a powder metallurgy process. The effects of the porosity and Zn content on
Accepted 23 January 2013 the microstructure and the mechanical properties of the fabricated scaffolds were studied.
Available online 8 February 2013 The above mentioned fabrication process involved sequential stages of mixing and

Keywords: compression of Mg and Zn powders with carbamide materials as space-holder particles

Magnesium followed by sintering the green compacts at different temperatures below the melting

Zinc point of Mg. The results indicate that the porous Mg–Zn specimens with a porosity and

Porous materials pore size of approximately 21–36% and 150–400 mm, respectively, could have enhanced

Porosity mechanical properties comparable with those of cancellous bone. In addition, an increase

Mechanical properties in the amount of Zn in the applied alloy gives rise to a significant refinement of

Sintering magnesium grain size and an improvement in the mechanical properties, such as the
compression strength, of the porous Mg–Zn specimens. Furthermore, according to the
results, the porous Mg–Zn alloy could be considered one of the most promising scaffold
materials for hard tissue regeneration.
& 2013 Elsevier Ltd. All rights reserved.

1. Introduction been investigated for tissue engineering of bone (Wen et al.,


2001; Zhuang et al., 2008). The major limitation of these
In recent decades, great interest in porous scaffold substi- porous materials is, however, their inadequate mechanical
tutes has stemmed from their applicability in bone tissue properties (Zhuang et al., 2008). For instance, the brittle
engineering applications (Zhuang et al., 2008). The scaffold nature of the porous bioactive ceramics or the very low
provides the necessary support for cells to proliferate and strength and Young’s modulus of the porous polymers –
maintain their differentiated function, and its architecture mainly lower than those of real human bones – have severely
defines the ultimate shape of the new bone. The ideal bone limited the applicability of these materials in load-bearing
substitute material should be osteoconductive, biodegrad- applications (Wen et al., 2004). Therefore, it is crucial to
able, and strong enough to fulfil the required load-bearing develop new scaffold materials with both a porous structure
functions. Several scaffold materials, including hydroxyapa- similar to natural bones and excellent mechanical properties.
tite (HA) and natural polymers (e.g., collagen and chitin), have Furthermore, scaffold materials that undergo a complete

n
Corresponding author. Tel.: þ98 21 77459151; fax: þ98 21 77240480.
E-mail address: zahraseyedraoufi@iust.ac.ir (Z.S. Seyedraoufi).

1751-6161/$ - see front matter & 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.jmbbm.2013.01.023
2 journal of the mechanical behavior of biomedical materials 21 (2013) 1 –8

biodegradation process after a proper time span in a human properties of the specimens with respect to different physical
body are even more desirable (Wen et al., 2004; Zhuang et al., properties. Until now, there has been no study on either the
2008). production of porous Mg–Zn scaffolds or the ways in which
Stainless steels, cobalt-based and titanium alloys have the mechanical properties and corrosion resistance of the
been widely studied and applied clinically in hard tissue magnesium scaffolds may be improved. In the present work,
implants, specifically in load-bearing applications, due to porous Mg–4 wt% Zn and Mg–6 wt% Zn specimens have been
their high strength, ductility and good corrosion resistance fabricated using a powder metallurgy process. The microstruc-
(Ng et al., 2009; Wanga et al., 2009; Yang et al., 2009). However, tures and the mechanical properties of the specimens contain-
they cannot degrade spontaneously, and a second surgical ing different zinc amounts and porosities are investigated.
procedure is usually needed to remove metal implants from
the body after the tissues have completely healed (Ng et al.,
2009; Wanga et al., 2009; Yang et al., 2009). In addition, after 2. Materials and methods
the healing stage, the presence of a permanent bone plate
has the potential to result in a number of adverse effects, Pure magnesium (purity Z99%, particle size r0.1 mm) and
most seriously the occurrence of osteoporosis in the neigh- pure zinc powder (purity Z99.8%, particle size r45 mm)
bouring bone tissues due to a mismatch in elastic modulus, purchased from Merck were utilised as the starting materials.
creating stress shielding (Ng et al., 2009; Wanga et al., 2009; Carbamide (CO(NH2)2) particles purchased from Merck were
Ye et al., 2010). For young patients, permanent bone plates employed as the space-holder agent. The particle size of the
also restrict bone growth. Repeated surgeries increase both spacer agent material was in the range of 200–400 mm with a
costs and patient morbidity (Xin et al., 2008). Magnesium has purity of 99.9%. After mixing the starting materials with the
been recently recognised as a very promising biomaterial for space-holder particles, porous Mg–4 wt% Zn and Mg–6 wt%
bone implants because it has excellent mechanical properties Zn specimens were prepared through a powder metallurgy
and is biodegradable and bioresorbable (Khanra et al., 2010; process. The mixtures of magnesium and zinc powders were
Kirkland et al., 2012; Li et al., 2008; Müller et al., 2007; Wen prepared based on two different zinc amounts, 4 and 6 wt%,
et al., 2001, 2004; Xin et al., 2008, 2011; Zhuang et al., 2008). while the carbamide particles were thoroughly added to the
Magnesium is the fourth most abundant cation in the human above specimens with variable volume contents of 15%, 25%
body (1 mol/adult) and is found mainly in bone tissue, and and 35%. Details of the fabricated samples are listed in
stabilising DNA and RNA structures (Khanra et al., 2010). Table 1.
Magnesium alloys enjoy low densities and high strength-to- The mixed powders were uniaxially pressed at a pressure
weight ratios (Krawiec et al., 2011; Wen et al., 2004). Porous of 100 MPa into green compacts with dimensions of 10 mm in
magnesium has the potential to serve as a degradable scaf- diameter and 15 mm in length. Differential thermal analysis
fold for bone substitute applications (Atrens et al., 2011; Wen (DTA) and thermal gravimetry analysis (TGA) (DTA–TGA:
et al., 2004; Zhuang et al., 2008). Some studies have shown Shimadzu, TGA-50, Japan) were conducted under an argon
that the dissolved magnesium ions may promote bone cell atmosphere to determine the temperatures at which the
attachment and tissue growth on the implants (Xin et al., carbamide particles were removed from the samples and
2008). when the reactions between Mg and Zn occurred during the
However, as a biomedical material, the high corrosion rate sintering process. The green compacts were subsequently
of magnesium alloys has been observed to result in high heat treated to burn out the spacer particles and to sinter the
concentrations of Mg ion and hydrogen gas release. It has porous Mg–Zn specimens separately in a tube furnace under
been found that different Mg alloys, such as AZ91, AZ31, an argon atmosphere. The heat treatment process consisted
AZ31, WE43 and LAE442, might be prone to localised corro- of two steps: first, heating up to 250 1C at a rate of 3.751/min
sion in in vitro and in vivo tests (Khanra et al., 2010). In and staying at that temperature for 4 h, and then heating up
addition, because the yield strength of pure magnesium is to 500, 550, 565 and 580 1C at a rate of 81/min and staying at
approximately 20 MPa, much lower than that of human long the final temperature for 2 h. The sintering temperature was
bone (106–133 MPa), it is essential to take relevant measures set slightly above the temperature at which DTA–TGA peaks
to control its corrosion rate and to improve its mechanical appeared, and after the sintering stage, the specimens were
properties. Zinc not only improves the corrosion resistance furnace cooled down to room temperature.
but also enhances the mechanical properties of magnesium Pore sizes and morphologies of the porous magnesium
alloys (Gu et al., 2010; Kaya et al., 2006; Yin et al., 2008; Zhang specimens were observed using optical microscopy. Total
et al., 2009). Moreover, zinc is recognised as a highly essential
element for humans (Zhang et al., 2009). It has been shown
Table 1 – Details of the fabricated samples.
that Mg–Zn alloys possess the highest capacity for precipita-
tion hardening due to segregation of an intermediate phase Zn (wt%) Carbamide (vol%) Sample
that ensures a combination of high strength and ductility
(Kaya et al., 2006). Gu et al. (2010) showed that Zn may 4 15 A1
4 25 A2
improve the plasticity of a Mg–Zn–Ca alloy, and Yin et al.
4 35 A3
(2008) reported that the addition of Zn can effectively 6 15 B1
increase the strength of a Mg–Zn–Mn alloy. 6 25 B2
Wen et al. (2001) and Zhuang et al. (2008) produced porous 6 35 B3
magnesium scaffolds and investigated the mechanical
journal of the mechanical behavior of biomedical materials 21 (2013) 1 –8 3

porosity (P) of the porous specimens was measured using testing of samples with dimensions of F10 mm  15 mm. The
gravimetry according to the following equation (Zhuang et al., tests were performed with a SANTAM (STM-20, Iran) testing
2008): machine at room temperature at a rate of 0.3 mm/s. Each
result was taken as the mean value of testing on five samples.
P ¼ ð1r=rs Þ  100% ð1Þ

where rs is the density of the magnesium–zinc specimen


evaluated via the immersion method and r is the apparent
3. Results and discussion
density of the porous magnesium–zinc specimen, which can be
measured by the weight-to-volume ratio of the porous specimen.
3.1. DTA–TGA analysis
Microstructural and elemental analyses were conducted
Fig. 1 illustrates the DTA–TGA curves of the A2 samples.
using scanning electron microscopy (SEM: TSCAN-VEGA,
As can be observed, the DTA analysis demonstrated an
China) and energy dispersive X-ray spectroscopy (EDS:
endothermic peak at E140 1C, which is accompanied by a
TSCAN-VEGA, China), respectively. An X-ray diffraction tech-
8.38% mass decrement.
nique (XRD: JDX-8030, Jeol, Japan) with a Cu-Ka source was
Such a decrease in the mass is interpreted as the removal
used to obtain diffraction patterns of the sintered samples.
of the spacer particles from the specimen. Furthermore, the
The compressive strength of the porous Mg–4 wt% Zn and
6.98% mass decrement from E500 1C to E580 1C indicates the
Mg–6 wt% Zn specimens was measured using compression
occurrence of the sintering process in this temperature range.
According to a phase diagram of the Mg–Zn binary alloy
system (Deng et al., 2005) and considering the data repre-
sented by the DTA–TGA curves, in order to achieve the
optimum sintering temperature, consolidation of the speci-
mens was initiated at variable sintering temperatures of 500,
550, 565 and 580 1C under an argon atmosphere.

3.2. Microstructure of porous specimens

Using Eq. (1), the porosity volume fractions of the three classes
of A1 (B1), A2 (B2) and A3 (B3) specimens are E21, E29 and
Fig. 1 – DTA/TGA curves of the Mg–4 wt% Zn powder E36 vol%, respectively. Fig. 2 shows the optical micrographs of
containing 25 vol% carbamide. the A1, A2 and A3 specimens, sintered at 550 1C. It is clear that

Fig. 2 – Optical micrographs of the A1, A2 and A3 specimens, sintered at 550 1C, with different porosity amounts: (a) 21%,
(b) 29% and (c) 36%.
4 journal of the mechanical behavior of biomedical materials 21 (2013) 1 –8

the pore sizes of the A1, A2 and A3 specimens are smaller than the grain growth effect would result in an obvious degrada-
200 mm, 100–300 mm and 200–400 mm, respectively. Visual char- tion of the mechanical properties.
acterisation of the pores revealed that two types of pores, open Fig. 4 shows the optical microstructures of the A1 and B1
interconnected macropores and small isolated micropores, samples sintered at 550 1C. As shown in the figure, increasing
were distinguishable within the microstructure of the porous the Zn content of the synthesised alloys produces a signifi-
specimens. Some of the micropores are formed as a result of cant reduction in the grain size of the matrix. Such an
volume shrinkage of the magnesium and zinc powders during observation indicates that the addition of Zn could effectively
the sintering process. In addition, the number of isolated pores refine grain size and improve mechanical properties.
increases with a decrease in the volume fraction of the micro-
structural porosities. 3.3. XRD analysis
Fig. 3 shows the optical micrographs of the A2 samples
sintered at 500, 550, 565 and 580 1C. As observed in the figure, X-ray diffraction patterns of the A2 samples sintered at
an increase in the sintering temperature leads to an incre- different temperatures are shown in Fig. 5. As observed in
ment in the grain size. This result implies that 550 1C should the patterns, Mg peaks were detected in all of the samples,
be considered the optimum sintering temperature because and MgO was detected as the minor phase in the patterns.

Fig. 3 – Optical micrographs of the A2 specimens sintered at different temperatures: (a) 500, (b) 550, (c) 565 and (d) 580 1C.

Fig. 4 – Optical micrographs of the porous Mg–Zn specimens sintered at 550 1C with different Zn amounts: (a) 4 wt% and
(b) 6 wt%.
journal of the mechanical behavior of biomedical materials 21 (2013) 1 –8 5

Fig. 6 – SEM micrographs of the A2 specimens sintered at


different temperatures: (a) 550 and (b) 580 1C.

Table 2 – EDS analysis results of the porous A2 samples.

Fig. 5 – XRD patterns of the A2 specimens sintered at the Temperature (1C) Mg content (at%) Zn content (at%)
temperatures of 500, 550, 565 and 580 1C.
550 18.73 19.91
580 16.67 43.04
In fact, the presence of the magnesium oxide peaks indicates
that MgO was most likely formed during the preparation and/
or sintering process; however, MgO is biologically a non-toxic The SEM results along with the Mg–Zn binary alloy phase
oxide (Kannan and Singh, 2008). The increased intensity of diagram (Deng et al., 2005) indicate that the Mg7Zn3 and
the MgO peaks is a consequence of sintering at higher MgZn phases can act as local cathodic agents under atmo-
temperatures and is a further evidence that 550 1C is the spheric conditions (Krawiec et al., 2011) and are formed at
optimum sintering temperature. However, XRD results could 550 1C and 580 1C, respectively. The XRD and SEM results
not reveal the presence of any other possible phases. indicated that the amount of precipitates (Mg7Zn3 and MgZn)
in the samples is very low because their corresponding peaks
3.4. SEM analysis could not be detected by XRD. EDS elemental mappings of the
Mg–4 wt% Zn samples with 21 vol% porosity and sintered at
SEM micrographs of the porous A2 samples sintered at 550 550 1C are shown in Fig. 7. From the figure, it can be inferred
and 580 1C are illustrated in Fig. 6. According to the EDS that Zn is uniformly distributed in the Mg matrix. Micro-
analysis results (which are listed in Table 2) of points ‘‘A’’ and structures of the samples, sintered at 550 and 580 1C (Fig. 6),
‘‘B’’ (marked with an  in Fig. 6), precipitates detected in the show that the grain size of the matrix increased at the higher
grain boundaries of the samples sintered at 550 and 580 1C sintering temperature, reinforcing that 550 1C could be the
might be Mg7Zn3 and MgZn, respectively. optimum sintering temperature, as previously mentioned.
6 journal of the mechanical behavior of biomedical materials 21 (2013) 1 –8

Fig. 7 – (a) SEM micrograph of the A1 sample sintered at 550 1C, (b) and (c) elemental mapping images of (a).

Fig. 8 – Compressive strength of the porous Mg–Zn specimens sintered at different temperatures with different Zn amounts:
(a) 4 wt% and (b) 6 wt%.

3.5. Mechanical properties What is evident in the figures is that the values of the
compressive strength and Young’s modulus for the porous
Figs. 8 and 9 illustrate the compressive strength and Young’s Mg–6 wt% Zn samples are higher than that for Mg–4 wt% Zn.
modulus values of the porous Mg–4 wt% Zn and Mg–6 wt% Zn The other conclusion that can be drawn from these results is
samples as a function of porosity at different sintering the remarkable grain size decrement in the specimens as a
temperatures. The compressive strength and Young’s mod- result of an increase in the Zn content. On the other hand, Zn
ulus of the porous samples decreased with an increase in the mainly resolves in primary magnesium when the Zn content
volume fraction of porosities at all sintering temperatures. is 1–2 wt%, which can improve the strength of the alloy by
journal of the mechanical behavior of biomedical materials 21 (2013) 1 –8 7

Fig. 9 – Young’s modulus of the porous Mg–Zn specimens sintered at different temperatures with different Zn amounts:
(a) 4 wt% and (b) 6 wt%.

Table 3 – Summary of mechanical properties of various porous materials and natural bone.

Porous materials Porosity (%) Pore size (mm) Flexure (F) and compressive (C) Modulus (GPa)
strength (MPa)

Porous magnesium 36–55 200–400 14.1–27 (F), 15–31 (C) 3.6–18.1


Porous polymers 58–80 300 12.4–56.2 (F), 2.7–11 (C) 0.05–1.2
Porous HA 50–77 200–400 2.1–7.2 (F), 1.2–17.4 (C) 0.12–7
Porous composite bioglass 77–80 100 1.5–3.9 (F), 0.42 (C) 0.14–0.26
(20–50 wt þPLLA)
Natural bone – – 1.8–150 (F), 2–180 (C) 0.1–20

solid-solution strengthening (Yin et al., 2008). However, when properties of the porous specimens with porosities of 19–36%
the Zn content is 4 or 6 wt%, such as in the specimens of the and with pore sizes of approximately 150–400 mm were
current research, many Mg–Zn phases would most likely be investigated by compressive testing. The results indicate that
precipitated from the matrix. Such an effect is expected to the mechanical properties (compressive strength and Young’s
enhance the strength by a dispersion strengthening mechan- modulus) of the porous Mg–Zn specimens improve with a
ism (Yin et al., 2008). Therefore, Zn can act as a constituent to reduction in the volume fraction of porosities. Moreover,
refine the microstructure of the matrix and to improve the increasing the amount of Zn could refine the magnesium
mechanical strength. In addition, it is shown that the highest grain size and, as a result, could improve the mechanical
compressive strength and Young’s modulus values occurred properties of the porous Mg–Zn alloy. However, compared to
in the specimens sintered at 550 1C. Therefore, these mechan- porous bioactive ceramic and polymeric scaffolds, the porous
ical property results are more evidence that confirms that Mg–Zn specimens have more appropriate mechanical proper-
550 1C is the optimum sintering temperature. The major ties that are closer to those of natural bone. Therefore, the
problems of the implant materials in use today are related to porous Mg–Zn metals have the potential to serve as degradable
the high Young’s modulus of metallic materials and the low implants for bone substitute applications.
mechanical properties of highly porous ceramics and poly-
mers. These problems can be solved by selecting a porous
Mg–Zn scaffold with the appropriate porosity. Although the
r e f e r e nc e s
compressive strength and Young’s modulus of the porous Mg–
Zn specimens decrease as porosity increases, the mechanical
properties of these specimens are closer to those of natural
Atrens, A., Liu, M., Abidin, N.I.Z., 2011. Corrosion mechanism
bone (Table 3; Zhuang et al., 2008). Therefore, it is suggested
applicable to biodegradable magnesium implants. Journal of
that porous Mg–Zn is a promising scaffold material for hard Materials Science and Engineering B 176, 1–28.
tissue regeneration. Deng, C.J., Wong, M.L., Ho, M.W., Yu, P., Ng, Dickon H.L., 2005.
Formation of MgO and Mg–Zn intermetallics in an Mg-based
composite by in situ reactions. Composites Part A 36, 551–557.
4. Conclusions Gu, X., Zheng, Y., Zhong, S., Xi, T., 2010. Corrosion of, and cellular
responses to Mg–Zn–Ca bulk metallic glasses. Biomaterials 31,
1093–1103.
Porous Mg–4 wt% Zn and Mg–6 wt% Zn specimens have been Kannan, M.B., Singh, R.R.K., 2008. Evaluating the stress corrosion
fabricated by a powder metallurgy process. According to the cracking susceptibility of Mg–Al–Zn alloy in modified-
results, the optimum sintering temperature to synthesise simulated body fluid for orthopaedic implant application.
the porous Mg–Zn specimens is E550 1C. The mechanical Scripta Materialia 59, 175–178.
8 journal of the mechanical behavior of biomedical materials 21 (2013) 1 –8

Kaya, A., Eliezer, D., Ben-Hamu, G., Golan, O., Na, Y.G., Shin, K.S., Wen, C.E., Yamada, Y., Shimojima, K., Chino, Y., Hosokawa, H.,
2006. Microstructure and corrosion resistance of alloys of the Mabuchi, M., 2004. Compressibility of porous magnesium
Mg–Zn–Ag system. Metal Science and Heat Treatment 48, foam: dependency on porosity and pore size. Materials Letters
50–54. 58, 357–360.
Khanra, A.K., Jung, H.C., Yu, S.H., Hong, K.S., Shin, K.S., 2010. Xin, Y., Huo, K., Tao, H., Tang, G., Chu, P.K., 2008. Influence of
Microstructure and mechanical properties of Mg–HAP aggressive ions on the degradation behavior of biomedical
composites. Journal of Materials Science 33 (1), 43–47. magnesium alloy in physiological environment. Acta
Kirkland, N.T., Birbilis, N., Staiger, M.P., 2012. Assessing the Biomaterialia 4, 2008–2015.
corrosion of biodegradable magnesium implants: a critical Xin, Y., Hu, T., Chu, P.K., 2011. Degradation behaviour of pure
review of current methodologies and their limitations. Acta magnesium in simulated body fluids with different
Biomaterialia 8, 925–936. concentrations of HCO3. Corrosion Science 53, 1522–1528.
Krawiec, H., Stanek, S., Vignal, V., Lelito, J., Suchy, J.S., 2011. The Yang, J.X., Cui, F.Z., Yin, Q.S., Zhang, Y., Zhang, T., Wang, X.M.,
use of microcapillary techniques to study the corrosion 2009. Characterization and degradation study of calcium
resistance of AZ91 magnesium alloy at the microscale. phosphate coating on magnesium alloy bone implant in vitro.
Corrosion Science 53, 3108–3113. IEEE Transactions on Plasma Science 37, 1161–1168.
Li, Z., Gu, X., Lou, S., Zheng, Y., 2008. The development of binary Ye, X., Chen, M., Yang, M., Wei, J., Liu, D., 2010. In vitro corrosion
Mg–Ca alloys for use as biodegradable materials within bone. resistance and cytocompatibility of nano-hydroxyapatite
Biomaterials 29, 1329–1344. reinforced Mg–Zn–Zr composites. Journal of Materials Science:
Müller, W.D., Nascimento, M.L., Zeddies, M., Córsico, M., 2007. Materials in Medicine 21, 1321–1328.
Magnesium and its alloys as degradable biomaterials. Yin, D., Zhang, E., Zeng, S., 2008. Effect of Zn on mechanical
Corrosion studies using potentiodynamic and EIS property and corrosion property of extruded Mg–Zn–Mn alloy.
electrochemical techniques. Journal of Materials Research 10 Transactions of Nonferrous Metals Society of China 18,
(1), 5–10. 763–768.
Ng, W.F., Wong, M.H., Cheng, F.T., 2009. Cerium-based coating for Zhang, E., Yin, D., Xu, L., Yang, L., Yang, K., 2009. Microstructure,
enhancing the corrosion resistance of bio-degradable Mg mechanical and corrosion properties and biocompatibility of
implants. JMaterials Chemistry and Physics 119, 1–5. Mg–Zn–Mn alloys for biomedical application. Materials
Wanga, Y.M., Wang, F.H., Xu, M.J., Zhao, B., Guo, L.X., Ouyang, J.H., Science and Engineering C 29, 987–993.
2009. Microstructure and corrosion behavior of coated AZ91 Zhuang, H., Han, Y., Feng, A., 2008. Preparation, mechanical
alloy by microarc oxidation for biomedical application. properties and in vitro biodegradation of porous magnesium
Applied Surface Science 255, 9124–9131. scaffolds. Materials Science and Engineering C 28, 1462–1466.
Wen, C.E., Mabuchi, M., Yamada, Y., Shimojima, K., Chiho, Y.,
Asahina, T., 2001. Processing of biocompatible porous Ti and
Mg. Scripta Materialia 45, 1147–1153.

Вам также может понравиться