You are on page 1of 8

Applied Surface Science 257 (2011) 74477454

Contents lists available at ScienceDirect

Applied Surface Science


journal homepage: www.elsevier.com/locate/apsusc

Microstructure and mechanical properties of selective laser melted magnesium


C.C. Ng a , M.M. Savalani a, , M.L. Lau b , H.C. Man a
a b

Department of Industrial and Systems Engineering, The Hong Kong Polytechnic University, Hong Kong Department of Mechanical Engineering, The Hong Kong Polytechnic University, Hong Kong

a r t i c l e

i n f o

a b s t r a c t
The effects of laser processing parameters on the microstructure and mechanical properties of selective laser-melted magnesium were investigated. The results show that the microstructure characteristics of the laser-melted samples are dependent on the grain size of SLM magnesium. The grains in the molten zone coarsen as the laser energy density increases. In addition, the average hardness values of the molten zone decreases signicantly with an increase of the laser energy densities and then decreased slowly at a relatively high laser energy density irrespective of mode of irradiation. The hardness value was obtained from 0.59 to 0.95 GPa and corresponding elastic modulus ranging from 27 to 33 GPa. The present selective laser-melted magnesium parts are promising for biomedical applications since the mechanical properties are more closely matched with human bone than other metallic biomaterials. 2011 Elsevier B.V. All rights reserved.

Article history: Received 12 November 2010 Received in revised form 27 February 2011 Accepted 1 March 2011 Available online 8 March 2011 Keywords: Magnesium Selective laser melting Microstructure Mechanical properties Hardness

1. Introduction During the last several decades, development of medical implant materials for use in the orthopedic eld has increased signicantly [13]. Metal materials are of considerable importance among diverse material types for load-bearing orthopedic applications [4,5]. Metallic biomaterials such as titanium, stainless steel, chromiumcobalt and its alloy are widely used due to their superior mechanical properties, biocompatibility, biodegradability and chemical stability [6]. However, the major disadvantages in expanding their use in the orthopedic sectors are the mismatch of mechanical properties between these biomaterials and natural bones, especially in the elastic modulus, which results in implants loosening following stress shielding of the bone [7] and also the release of toxic species which may cause allergy and cancer [8]. One method to alleviate these problems is to investigate a new kind of metal-based biocompatible material with appropriate mechanical properties close to that of human bone and with an excellent biocompatibility as well as low toxicity. Recently, magnesium has been suggested as a very promising metallic biomaterial which offers both inherent biocompatible performance [9,10] and favorable mechanical properties in relation to the natural bone compared with any other widely used metallic biomaterials [11]. Therefore, stress shielding effects can be minimized by the use of magnesium implant. In addition, its ability to degrade in situ also implies that the need for secondary surgery whole elim-

Corresponding author. Tel.: +852 3400 3190; fax: +852 2362 5267. E-mail address: mmfsmm@inet.polyu.edu.hk (M.M. Savalani). 0169-4332/$ see front matter 2011 Elsevier B.V. All rights reserved. doi:10.1016/j.apsusc.2011.03.004

inated due to gradually dissolvable, resorption and excretion of the magnesium implant in human body environment [12]. Currently, substantial efforts have been put to investigate the potential of magnesium to replace presently used metallic biomaterials for orthopedic applications [13]. Selective laser melting (SLM) is a new approach of manufacturing parts directly from metal powders [14]. Among those process parameters, laser power and laser scan speed were found to have a direct inuence on the microstructures and mechanical properties of the laser-melted parts [15]. The relationship between the parameters can be combined into a term called linear energy density. The linear energy density governs the amount of laser power incident to the powder bed per unit area per unit scan speed. Many researchers showed that the improvement of the surface mechanical properties by laser rapid solidication processing could be obtained by varying the process parameters and is related to the rened microstructures in the molten zone. Zheng et al. [16] illustrated that laser deposited material experiences a signicant rapid quenching effect and a very high cooling rate can be attained, where the microstructural evolution is related to process parameters during the laser engineered net shaping (LENS ) process, which is a laser-powder additive manufacturing technology for fabricating metal parts directly from a computer-aided design (CAD) solid model by using a metal powder injected into a molten pool created by a focused, high-powered laser beam. High cooling rates were produced by high power density welding, resulting in microstructures that are far from equilibrium and the cooling rates at the solid-liquid interface were found to be from 104 to 105 K/s [17]. Laser surface alloying under pulsed-wave irradiation resulted in even higher cooling rates. Samant and Dahotre [18] illustrated that

7448

C.C. Ng et al. / Applied Surface Science 257 (2011) 74477454

extremely high cooling rates (the order of 108 K/s) were obtained in the laser deposition of Al + Al2 O3 on magnesium alloys processed using a pulsed Nd:YAG laser. In the presence of a high cooling rate, grain structures do not grow signicantly during solidication [19]. Effects of process parameters on the microstructural characteristics of laser surface melted AISI 410 stainless steel have been studied by Krishna and Bandyopadhyay [20] and it was found that grain renement after surface laser melting lead to an increase in surface hardness. The fusion zone with the rened microstructures was found to be down to a maximum depth of 135 10 m. Yu et al. [21] demonstrated the correlation between yield strength of the weld joint and heat input of ber laser welding of Mg alloy and showed that yield strength increases with decreasing welding heat input due to ner microstructure obtained. Electron beam welding of MgAl based alloy has been conducted by Su et al. [22] and presented that the grains inside the fusion zone were nearly equixed in sharp with about 10 m in size, due to rapid cooling rate and the renement of grains responsible for the increase in microhardness in the fusion zone. The nanoindentation technique has recently been used in the biomedical engineering sector to measure mechanical properties of human bones as well as metallic implants and other biomaterials [23,24]. In a nanoindentation instrument, hardness is recorded progressively by measuring the loads and displacements prole when the indenter penetrates the surface of the material under test. The hardness values and elastic modulus information of the material can be determined by analyzing the loaddisplacement characteristics [25]. One of the main advantages of nanoindentation method is the capability of measuring a materials properties on an extremely small scale. At present, the nanoindentation technique has been introduced to measure the mechanical properties of the laser-melted magnesium tracks. To our knowledge, no previous study is concerned with the mechanical properties of SLM magnesium. Therefore, in this paper, the effects of laser process parameters on the microstructure (grain size and morphology of the -Mg) in the molten zone of SLM magnesium and the mechanical properties (hardness and elastic modulus) were investigated. 2. Methods 2.1. Processing SLM of pure magnesium powder was conducted on a GSI Lumonics Marker SPe Nd:YAG ( = 1.06 m) laser. Single tracks of 20 mm long were fabricated to investigate the effects of laser process parameters on the microstructure characteristics and mechanical properties. The detailed experimental setup and procedures have been described in the authors previous study [26]. 2.2. Microstructural analysis To reveal the microstrucutral characteristic of the laser-melted magnesium, the longitudinal cross-sectional surface of the samples were grounded and polished under water on Silicon carbide grinding paper from lower grade 60 up to higher grade 2400. In order to obtain good grinding surface, the abrasion direction of specimen was perpendicular to the pre-existing scratches. The specimen was polished by gamma diamond paste up to grade 1 um until scratchfree surface was obtained and etched with picric acid solution (6 g picric acid + 100 ml ethanol + 1 ml H3 PO4 ) according to ASTM Standard E407-07 [27]. Metallurgical observation of each sample were performed by means of a secondary scanning electron imaging technique (JEOL scanning electron microscope), operated at an accelerating voltage of 20 kV. The grain boundaries were marked manually and computed statistically to obtain the average grain

size of each sample. 30 intercepts were for each measurement were taken by linear line intercept method, based on ASTM Standard E112-96 [28]. 2.3. Indentation testing Nanoindentation tests were carried out using a Triboindenter (Hysitron, Inc.) equipped with a diamond Berkovich probe. After the preliminary trails to obtain the best loading and unloading parameters, the measurement sequence is as follows. The indentation load was xed at 8 mN at room temperature and the load and unload rates were kept at a constant of 0.8 mN. After the maximum load was held for 5 s, the samples were unloaded to 10% of the maximum load and kept for 15 s for thermal drift correction. The values of the nanomechanical properties of the molten zone of the longitudinal cross-section were made from an average value of 60 data points. The optical microscope was used to examine the surface topographies prior to and after indentation. The reduced elastic modulus (Er) and hardness (nHV) were computed from the load-displacement curve [23]. 3. Results and discussion 3.1. Microstructure Longitudinal cross-section microstructures of the conventional cast pure magnesium ingot and the selective laser-melted magnesium are presented in Fig. 1. A typical optical micrograph of the conventional casting has an equiaxed grain structure with the average grain size of about 7 m determined by linear line intercept method. The grain size is relatively large as compared with the laser-melted specimen due to the slower cooling rates. It can be inferred that the microstructure of the laser-melted magnesium with a higher solidication rate is much ner (25 m) than the conventional cast magnesium (Fig. 1b). The result is similar to the previous study by Teng et al. which showed that a highly rened microstructure of magnesium alloy was obtained under rapidly solidied process compared to the cast magnesium alloy due to the rapid solidication during processing [29]. 3.1.1. The typical microstructure of the continuous wave and pulsed wave laser melted magnesium Under continuous wave irradiation, the center of the lasermelted track had a ne, uniform and single phase structure with relatively small grains as shown in Fig. 2a. The -Mg single phase solidied in form of equiaxed crystals. The laser melting leads to the formation of fully recrystallised grains in melted zones of the lasermelted track. This microstructure has also been observed by other laser processing techniques of magnesium alloy such as laser welding [17,30]. Since the peak temperature of the melted zone is higher than the recrystallisation temperature and a longer time of laser irradiation in CW mode as compared with pulse mode, complete recrystallisation and complete grain growth occurred in the melted zone after laser melting [31]. This is a possible reason why the melted zone consists of ne homogeneous and full recrystallised -Mg grains. Unlike the continuous wave irradiation, laser-melted tracks fabricated under pulsed irradiation consist of incomplete growth of the -Mg phase (Fig. 2b). Since the solidication rate achieved under pulsed laser irradiation is higher than continuous wave irradiation [32], this inhibits the full growth of -Mg grains and since the interaction time is short, it has insufcient time for the crystals to arrange themselves such that thermodynamic equilibrium prevails at the solid/liquid interface, the average size of the grains is thus smaller than those grains obtained in the continuous

C.C. Ng et al. / Applied Surface Science 257 (2011) 74477454

7449

Fig. 1. Optical micrographs of (a) cast pure Mg ingot and (b) laser-melted Mg at longitinudual cross-section.

Fig. 2. Typical microstrucutral images of laser-melted magnesium under, (a) continuous wave irradiation at 1.27 109 J/m2 , (b) pulsed wave irradiation at 1.13 1012 J/m2 .

mode laser melted tracks, as determined by the linear intercept method. 3.1.2. Effects of the laser processing parameters on the microstructures of CW samples Under continuous wave irradiation, in order to discern grain size of the laser melted tracks processed under different laser energy densities, scanning electron microscopy (SEM) was preformed for further observation. Acceptable melting of magnesium takes place between the energy densities values of 1.27 109 J/m2 and 7.84 109 J/m2 . When the energy density values are below 1.27 109 J/m2 , the single line specimens could not be melted or were too fragile to handle [25]. This may possibly be because the melting temperature of magnesium (650 C) has not been reached. Higher energy values were not attainable due to the limitation of the laser. Thus, SEM images of the melted zones of the 1.27 109 J/m2 , 2.11 109 J/m2 , 3.92 109 J/m2 , 6.33 109 J/m2 and 7.84 109 J/m2 samples are shown in Fig. 3. The laser-melted magnesium is free of pores or cracks. However, it can be seen that the laser-melted magnesium has some white phases along the grain boundaries of -Mg grains. The grain boundary precipitates referred to magnesium oxide according to EDX analysis, and it is associated with oxygen pick-up during SLM process. The rapid solidication rates achieved after laser melting inhibit the partitioning of the oxide elements between the -Mg grains due to too little time for the MgO atoms to arrange themselves such that thermodynamic equilibrium prevails at the grain boundary [33]. As expected, various laser parameters resulted in different solidication rates of the molten pool and thermal cycles, leading to variation in microstructures of the melted zone. The SEM images conrm that all the melted zones consist of only equiaxed -Mg single phase with a high density of intergrainular boundaries. The average grain

sizes are about 2.30 m, 2.82 m, 3.65 m 4.10 m and 4.87 m, respectively. Evidently, the average grain size of the -Mg grains is larger for higher laser energy density values, the increase in the average grain size with increasing laser energy density is expected, since the temperature during laser melting generally increases with the increasing laser energy density. Therefore, the melted zone shows coarser grain at a lower cooling rate. The lower the cooling rate during solidication, the longer the time available for grain coarsening [34]. Thus, increasing laser energy density would result in grain coarsening. This trend can explain the relationship between grain size and laser energy density in the melted zone of the present samples, as illustrated in Fig. 4a. Similarly, as shown in Figs. 4b and 5, the increase of the laser power or decrease of the laser scan speed lead to the coarsening of the grains in the melted zone because the higher the laser power or slower the laser scan speed provided more driving force for grain boundary movement which then promoted the growth of grains. Previous studies on laser welding of magnesium alloys also indicated that grain coarsening occurred by increasing the heat input and improved heat conduction [35]. 3.1.3. Effects of the laser processing parameters on the microstructures of pulsed samples Under pulsed wave irradiation, the energy density provided by pulsed wave irradiation for melting track fabrication is much higher, ranging between 1.13 1012 J/m2 and 9.80 1012 J/m2 due to the higher peak power under pulse wave irradiation. Despite energy densities as high as 1 1013 J/m2 insufcient pulse duration time (20 ns) may have not brought the molten material to the evaporation temperature (1050 C) since no vaporization was observed during processing. Energy densities below 1 1012 J/m2 have resulted in the lack of melting of powders. SEM micrographs (Fig. 7) of the equiaxed structures of the melted zone show a very

7450

C.C. Ng et al. / Applied Surface Science 257 (2011) 74477454

Fig. 3. Microstructural images of the continuous wave laser-melted zone with different laser energy densities, (a) 1.27 109 J/m2 , (b) 2.11 109 J/m2 , (c) 3.92 109 J/m2 , (d) 6.33 109 J/m2 and (e) 7.84 109 J/m2 . Table 1 The grain size, hardness and elastic modulus with various energy densities under continuous wave irradiation. Energy density (J/m2 ) Grain size (m) Hardness (GPa) Young modulus (GPa) 1.27 109 2.30 0.3 0.87 0.13 32.76 1.07 2.11 109 2.82 0.2 0.75 0.06 27.99 4.41 3.92 109 3.65 0.5 0.67 0.05 26.22 3.67 6.33 109 4.10 0.3 0.62 0.09 24.09 2.46 7.84 109 4.87 0.7 0.59 0.05 20.84 4.06

ne grain size (2.35 m). Like the continuous wave irradiation, white precipitates are present along the grain boundaries of the Mg grains. The grains of the melted zone in all different energy densities are uniformly distributed as shown in Fig. 6(ae), and the effect of laser energy density on the average grain size of the melted zone is given in Fig. 7a. It is shown that the average grain size increases from approximately 2.3 m to 5.0 m with the increase in laser energy density from 1.13 1012 J/m2 to 9.80 1012 J/m2 . The results above indicated that different degrees of grain renement can be obtained under different laser energy densities. This was attributed to the difference in solidication and cooling rates as observed in the continuous wave irradiation. Finer grains can be obtained by fast cooling whereas slow cooling leads to the formation of coarse grains. Likewise, the results for the quantitative analysis of the effect of the laser power and the laser scan speed on the average grain size of -Mg are shown in Figs. 7b and 8, respectively. When the laser power was low or the laser scan speed was high, the average gain size became small since decreasing the laser power in effect acts in the same way as reducing the preheat temperature of the powder when the next pulse hits the base. The relatively high cooling rate at low laser power or high laser scan speed restrained the growth of -Mg grain during solidication. Su et al. showed that high peak power density would lead in increasing

the absorption of energy in the material and a reduction in cooling rate at the center of weld pool during laser welding [22]. On the contrary, Mg grains in the melted zone coarsened with either an increase in laser power or a reduction in scan speed.

3.2. Hardness 3.2.1. Effects of the laser processing parameters on the mechanical properties of continuous wave samples Nanoindentation was performed on the longitudinal crosssection of the laser-melted tracks. Fig. 9 shows the hardness proles against different laser energy densities ranged from 0.59 GPa to 0.87 GPa. As the laser energy density increased, the melted zone hardness value decreased. Table 1 summarizes the results of the grain size, hardness and elastic modulus with samples fabricated at various energy densities under continuous wave irradiation. Hardness values of the melted zone were 0.87 GPa, 0.75 GPa, 0.67 GPa, 0.63 GPa, 0.62 GPa and 0.59 GPa for laser energy density values of 1.27 109 J/m2 , 2.11 109 J/m2 , 3.92 109 J/m2 , 6.33 109 J/m2 and 7.84 109 J/m2 samples, respectively. The higher hardness values fabricated with low laser energy densities are attributed to laser rapid solidication and subsequent grain renement [36].

C.C. Ng et al. / Applied Surface Science 257 (2011) 74477454 Table 2 The grain size, hardness and elastic modulus with various energy densities under pulsed wave irradiation. Energy Density (J/m2 ) Grain size (m) Hardness (GPa) Young modulus (GPa) 1.13 1012 2.33 0.09 0.95 0.08 38.18 4.04 1.63 1012 2.50 0.47 0.85 0.02 38.10 1.66 2.64 1012 3.3 0.35 0.80 0.02 37.20 0.31 3.27 1012 3.82 0.20 0.69 0.02 36.25 2.28

7451

9.80 1012 4.86 0.31 0.67 0.05 27.41 1.6

Table 3 Summary of the mechanical properties of various metallic implant materials in comparison to natural bone edited from Ref. [10]. Properties Hardness (GPa) Elastic modulus (GPa) Natural bone 0.390.77 320 SLM magnesium 0.590.95 2733 Ti alloy 1.422.67 110117 Stainless steel 1.371.77 189205

a
Aveage Grain Size (um)

Average Grain Size (um)

5 4 3 2 1 0 0 2x10
9

5 4 3 2 1 Laser Power = 21W 0 0 10 20 30 40 50 60 70

4x10

6x10

8x10
3

Linear Enery Density (J/m )

Scan Speed (mm/s)


Fig. 5. The effect of laser scan speed on the average grain size of -Mg under continuous wave irradiation.

b
Average Grain Size (um)

6 5 4 3 2 1 Scan Speed = 10 mm/s 0 0 10 20 30 40

Laser Power (W)


Fig. 4. (a) The effect of laser energy density on the average grain size of -Mg under continuous wave irradiation. (b) The effect of laser power on the average grain size of -Mg under continuous wave irradiation.

3.2.3. Effects of the grain size on the hardness of both continuous and pulsed wave samples The melted zones had homogeneous equiaxed grain structures, and the hardness values increased as the laser energy density decreased. The higher cooling rate induced by low laser energy density resulted in smaller grain sizes, so that the hardness values were mainly related to the grain size in the melted zone. Fig. 11 shows the relationship between the hardness (GPa) and the reciprocal of the square root of grain size (d) for the melted zone. The hardness value (GPa) is found to be proportional to the reciprocal of the square root of grain size (d). According to the HallPetch equation for the melted zone of the present sample is expressed as [38] H = H0 + kd1/2 (2)

3.2.2. Effects of the laser processing parameters on the mechanical properties of pulsed wave samples The effect of laser energy density on the hardness of the melted zone is shown in Fig. 10. It can be seen that the lower the laser energy density is, the higher is the hardness of the melted zone. The increase of the hardness values can be attributed to the grain renement effect [37], as observed in continuous wave mode. During laser melting process, grain coarsening occurred in melted zone with an increase of laser energy density. Hence, a higher hardness value increases with decreasing in laser energy density. Moreover, it can be seen that with a further increase of laser energy density up to 2.11 109 J/m2 , the hardness value changed slightly, the effect of grain coarsening on the decrease of the hardness is partially offsetted. Table 2 summarized the results of the grain size, hardness and elastic modulus with samples fabricated at various energy densities under pulsed wave irradiation.

where H is the hardness of material, H0 and k are appropriate constants associated with the hardness measurements. The regression equations for continuous-wave irradiation and pulsed-wave irradiation are found to be y (GPa) = 0.069x (m1/2 ) + 0.493 (GPa) and y (GPa) = 0.068x (m1/2 ) + 0.644 (GPa) respectively (as shown in Fig. 11). The smaller the grain size is, the higher is the hardness value. The results conrm that the HallPetch relationship persists down to at least the nest grain size examined experimentally (2 m). Hence, the change of the hardness value is inversely proportional to the square root of the grain size. This illustrated that grain coarsening in the melted zone due to the effect of the thermal cycling resulted in the reduction of the hardness of the lasermelted magnesium, which is consistent with the microstrucutres displayed above. It can be concluded that smaller -Mg size crystals resulted in higher hardness values. In the HallPetch relationship only high angle grain boundaries are considered as obstacles to the dislocation movement. The ner grain size can facilitate the grain boundary sliding process [25].

7452

C.C. Ng et al. / Applied Surface Science 257 (2011) 74477454

Fig. 6. SEM images of the microstructure of the pulsed wave laser-melted zone with different laser energy densities, (a) 1.13 1012 J/m2 , (b) 1.63 1012 J/m2 , (c) 2.64 1012 J/m2 , (d) 3.27 1012 J/m2 and (e) 9.80 1012 J/m2 .

a
Average Grain Size (um)

6 5 4 3 2 1 0

Average Grain Size (um)

5 4 3 2 1 Laser Power = 26W 0 0 10 20 30 40 50 60 70

2x10

12

4x10

12

6x10

12

8x10
3

12

1x10

13

Linear Enery Density (J/m )

Scan Speed (mm/s)


Fig. 8. The effect of scan speed on the average grain size of -Mg under pulsed wave irradiation.

b
Average Grain Size (um)

6 5 4 3 2 1 Scan Speed = 10 mm/s 0 0 10 20 30 40

1.2 1.0

Hardness (GPa) Laser Power (W)

0.8 0.6 0.4 0.2 0.0 0 2x109 4x109 6x109


3

8x109

1x1010

Fig. 7. (a) The effect of laser energy density on the average grain size of -Mg under pulsed wave irradiation. (b) The effect of laser power on the average grain size of -Mg under pulsed wave irradiation.

Linear Enery Density (J/m )


Fig. 9. The effect of laser energy density on the hardness of laser-melted magnesium under continuous wave irradiation.

C.C. Ng et al. / Applied Surface Science 257 (2011) 74477454

7453

1.2 1.0 0.8 0.6 0.4 0.2 0.0 0 2x1012 4x1012 6x1012
3

Hardness (GPa)

[10]. Compared with the conventional metallic biomaterials such as titanium alloy and stainless steel, the mechanical properties of the laser-melted magnesium show the closest to the natural bone. The remarkable reduction in hardness and elastic modulus of the lasermelted magnesium has a potential contribution towards using as a bone xation implant (Table 3). 4. Conclusions Microstructural characteristics and mechanical properties of SLM magnesium were investigated. The microstructure of molten zone consists of ne equiaxed grains of the -Mg phase after SLM process. The microstructure evolution was explained in terms of high cooling rate and rapid solidication during SLM. An increase of laser energy density resulted in a decrease of cooling rate and the grain coarsening of -Mg in the melted zone. The hardness value in the melted zone increased with decreasing grain size, which is in agreement with the HallPetch equation. An increase of laser energy density, the hardness value of melted zone decreased sharply at rst and then decreased slightly. The hardness values and elastic modulus of the laser melted magnesium were in the range of 0.59 and 0.95 GPa and 27 and 33 GPa, respectively, which is comparable with the cast magnesium ingot and are closely matched with the mechanical properties of human bone than other conventional metallic biomaterials. Acknowledgement
y = 0.069x + 0.493 2 R = 0.928

8x1012

1x1013

Linear Enery Density (J/m )


Fig. 10. The effect of laser energy density on the hardness of laser-melted magnesium under pulsed wave irradiation.

1.0

0.8

Hardness (GPa)

0.6

0.4

0.2

0.0 0.4

0.5

0.6
-1/2

0.7

The authors would like to express their gratitude for nancial support from the Hong Kong Polytechnic University, under the project code A-PK57 and the Department of Industrial Systems Engineering, The Hong Kong Polytechnic University. References
[1] P. Roach, D. Eglin, K. Rohde, C.C. Perry, Modern biomaterials: a review-bulk properties and implications of surface modications, J. Mater. Sci.: Mater. Med. 18 (2007) 12631277. [2] G. Balasundaram, T.J. Webster, A perspective on nanophase materials for orthopedic implant applications, J. Mater. Chem. 16 (2006) 37373745. [3] E.E. Stroganova, N.Y. Mikhailenko, O.A. Moroz, Glass-based biomaterials: present and future (a review), Glass Ceram. 60 (2003) 315319. [4] R.M. Pilliar, Porous-surfaced metallic implants for orthopedic applications, J. Biomed. Mater. Res.-A 21 (1987) 133. [5] C.E. Wen, Y. Yamada, K. Shimojima, Y. Chino, H. Hosokawa, M. Mabuchi, Novel titanium foam for bone tissue engineering, J. Mater. Res. 17 (2002) 26332639. [6] N.A. Al-Mobarak, Comparative study of some metallic biomaterials used as implants, Mater. Und Werkstofftechnik 39 (2008) 486491. [7] M. Geetha, A.K. Singh, R. Asokamani, A.K. Gogia, Ti based biomaterials, the ultimate choice for orthopaedic implantsa review, Prog. Mater. Sci. 54 (2009) 397425. [8] Y Okazaki, E. Gotoh, Comparison of metal release from various metallic biomaterials in vitro, Biomaterials 26 (2005) 1121. [9] U.Y. Qiao, J.C. Gao, Y. Wang, S. Wang, S. Wu, Y. Xue, Biocompatibility evaluation of magnesium-based materials: 2006 BIMW: 2006, Beijing Int. Mater. Week, Pts 14 546549 (2007) 459462. [10] M.P. Staiger, A.M. Pietak, J. Huadmai, G. Dias, Magnesium and its alloys as orthopedic biomaterials: a review, Biomaterials 27 (2006) 17281734. [11] K.Y. Chiu, M.H. Wong, F.T. Cheng, H.C. Man, Characterization and corrosion studies of uoride conversion coating on degradable Mg implants, Surf. Coat. Technol. 202 (2007) 590598. [12] G.L. Song, S.Z. Song, A possible biodegradable magnesium implant material, Adv. Eng. Mater. 9 (2007) 298302. [13] O. Duygulu, R.A. Kaya, G. Oktay, A.A. Kaya, Investigation on the potential of magnesium alloy AZ31 as a bone implant, 2006 BIMW: 2006, Beijing Int. Mater. Week, Pts 14 546549 (2007) 421424. [14] X. Yan, P. Gu, A review of rapid prototyping technologies and systems, Comp.Aid. Des. 28 (1996) 307318. [15] M.M. Savalani, L. Hao, R.A. Harris, Evaluation of CO2 and Nd: YAG lasers for the selective laser sintering of HAPEX (R), Proc. Inst. Mech. Eng. Part B: J. Eng. Manuf. 220 (2006) 171182. [16] B. Zheng, Y. Zhou, J.E. Smugeresky, J.M. Schoenung, E.J. Lavernia, Thermal behavior and microstructure evolution during laser deposition with laser-engineered net shaping: part II. Experimental investigation and discussion, Metall. Mater. Trans. A (Phys. Metall. Mater. Sci.) 39 (2008) 22372245.

(Grain Size)

(um

-1/2

b
Hardness (GPa)

1.0

0.8

0.6

0.4

0.2

y = 0.068x + 0.644 2 R = 0.931

0.0 0.4

0.5

0.6

0.7

(Grain Size)-1/2 (um-1/2)


Fig. 11. The relationship between hardness and grain size under (a) continuouswave irradiation and (b) pulsed-wave irradiation.

Table 3 depicts the range of elastic modulus and hardness values measured on the laser-melted magnesium using the nanoindentation technique, and compared with the mechanical properties of some commonly used metallic biomaterials [39]. It can be seen that the obtainable hardness values of laser-melted magnesium ranged between 0.59 and 0.95 GPa and the elastic modulus varied between 27 and 33 GPa, which are within the range of conventional cast magnesium (28 GPa). It is worth noting that the higher the value of elastic modulus of the laser-melted magnesium, the stronger the mismatch between the implant and living bone tissue, because of the small value of the elastic modulus of the bone tissues (320 GPa)

7454

C.C. Ng et al. / Applied Surface Science 257 (2011) 74477454 [29] H.T. Teng, X.L. Zhang, Z.T. Zhang, T.J. Li, S. Cockcroft, Research on microstructures of sub-rapidly solidied AZ61 magnesium alloy, Mater. Charact. 60 (2009) 482486. [30] X. Cao, M. Jahazi, J.P. Immarigeon, W. Wallace, A review of laser welding techniques for magnesium alloys, J. Mater. Process. Technol. 171 (2006) 188204. [31] B. Basu, A.W. Date, Rapid solidication following laser melting of pure metalsstudy of pool and solidication characteristics, Int. J. Heat Mass Transf. 35 (1992) 10591067. [32] F.M. Ghaini, M.J. Hamedi, M.J. Torkamany, J. Sabbaghzadeh, Weld metal microstructural characteristics in pulsed Nd: YAG laser welding, Scripta Mater. 56 (2007) 955958. [33] G. Abbas, Z. Liu, P. Skeldon, Corrosion behaviour of laser-melted magnesium alloys, Appl. Surf. Sci. 247 (2005) 347353. [34] V.T. Swamy, S. Ranganathan, K. Chattopadhyay, Resolidication behaviour of laser surface-melted metals and alloys, Surf. Coat. Technol. 71 (1995) 129134. [35] Y.J. Quan, Z.H. Chen, X.S. Gong, Z.H. Yu, Effects of heat input on microstructure and tensile properties of laser welded magnesium alloy AZ31, Mater. Charact. 59 (2008) 14911497. [36] D. Min, J. Shen, S.Q. Lai, J. Chen, Effect of heat input on the microstructure and mechanical properties of tungsten inert gas arc butt-welded AZ61 magnesium alloy plates, Mater. Charact. 60 (2009) 15831590. [37] M.-x. Ren, B.-s. Li, C. Yang, H.-z. Fu, Hardness and elastic modulus of microcastings by nanoindentation, Chin. J. Nonferrous Met. 18 (2008) 231236. [38] Y.-l. Gao, C.-s. Wang, H.-b. Liu, M. Yao, Effect of laser power on microstructure and properties of melted layer of AZ91HP magnesium alloy, J. Dalian Univ. Technol. 47 (2007) 469472. [39] A.M. Pietak, M.P. Staiger, J. Huadmai, G. Dias, Magnesium and its alloys as orthopedic biomaterials: a review, Biomaterials 27 (2006) 17281734.

[17] S. Lathabai, K.J. Barton, D. Harris, P.G. Lloyd, D.M. Viano, A. McLean, Welding and weldability of AZ31B by gas tungsten arc and laser beam welding processes, Minerals, Metals & Materials Soc., Warrendale, 2003. [18] A.N. Samant, N.B. Dahotre, Computational prediction of grain size during rapid laser surface modication of Al-O ceramic, Phys. Status Solidi-R 1 (2007) R4R6. [19] C. Kwakernaak, T.J. Nijdam, W.G. Sloof, Microstructure renement of NiCoCrAlY alloys by laser surface melting, Metall. Mater. Trans. A: Phys. Metall. Mater. Sci. 37A (2006) 695703. [20] B.V. Krishna, A. Bandyopadhyay, Surface modication of AISI 410 stainless steel using laser engineered net shaping (LENS), Mater. Des. 30 (2009) 14901496. [21] L. Yu, K. Nakata, N. Yamamoto, J. Liao, Texture and its effect on mechanical properties in ber laser weld of a ne-grained Mg alloy, Mater. Lett. 63 (2009) 870872. [22] S.F. Su, J.C. Huang, H.K. Lin, N.J. Ho, Electron-beam welding behavior in MgAlbased alloys, Metall. Mater. Trans. A 33A (2002) 14611473. [23] F. Haque, Application of nanoindentation to development of biomedical materials, Surf. Eng. 19 (2003) 255268. [24] D.M. Ebenstein, L.A. Pruitt, Nanoindentation of biological materials, Nano Today 1 (2006) 2633. [25] M.X. Zhang, H. Huang, K. Spencer, Y.N. Shi, Nanomechanics of MgAl intermetallic compounds, Surf. Coat. Technol. 204 (2010) 21182122. [26] C.C. Ng, M.M. Savalani, H.C. Man, I. Gibson, Layer manufacturing of magnesium and its alloy structures for future applications, Virtual Phys. Prototyping 5 (2010) 1319. [27] A.S. E407-07, Standard Practice for Microetching Metals and Alloys, ASTM Standards (2007). [28] A.S. E112-96, Standard Test Methods for Determining Average Grain Size, ASTM Standards (2004).

You might also like