Está en la página 1de 10

2009-01-0020

Effect of Microstructure on the properties of a biomedical Co-Cr-Mo Alloy Produced by MIM P. Vieira Muterlle *, M. Zendron**, C. Zanella*, M. Perina **, R. Bardini**, A. Molinari * * University of Trento (Italy) ** MIMEST spa, Pergine, Trento (Italy) Abstract In this work the influence of microstructure on the corrosion and wear properties of a Co-Cr-Mo alloy produced by Metal Injection Moulding, was investigated. The materials have been sintered to full density and heat treated to dissolve carbides, as well as to modify the constitution of the metallic matrix. The sintering cycle transforms the f.c.c. matrix of the powder in an almost fully h.c.p. one. On solution annealing, simultaneously to the either full or partial dissolution of carbides, h.c.p. reverts partially to f.c.c. The microstructural characterization was carried out by SEM, and by XRD the volume fractions of the matrix constituents were analyzed. The microstructural characteristics were then correlated to the results of corrosion and wear properties. Keywords: Co-29Cr-6Mo alloys, MIM, corrosion and wear properties. 1. Introduction The Co-29Cr-6Mo alloy used in biomedical field can be produced with different carbon contents, from 0.05% up to 0.35%. On increasing the carbon content, the sintering temperature can be decreased from 1380C down to 1300C to obtain the full density, because of the formation of an eutectic liquid which activates sintering [1]. Consequently, the as sintered microstructure may contain different kind of carbides, which have a great influence on the properties. Moreover, the formation of carbides promotes the precipitation of sigma phase [2, 3], which also effect properties. The effect of carbides and sigma phase on mechanical properties is well assessed: they increase hardness and yield strength, but decrease ductility [4, 5]. In some previous works, the effect of post sintering heat treatment, consisting of solution annealing and aging, on tensile properties of the Co-29Cr-6Mo produced by MIM of prealloyed powders with different carbon contents was investigated [6, 7, 8]. The influence on the wear resistance and the corrosion resistance of MIM products is not yet established. As far as wear resistance is concerned, a study on a cast alloy concludes that carbides may cause abrasive wear of the counterface polymeric material, which could release debris at the metal-polyethylene interface [9]. The UHMWPE wear debris is mainly responsible for periprosthetic tissue reaction, osteolysis, and eventually, late aseptic loosening. In a metal-on-metal wear tests, the effect of the constitution of the metallic matrix on wear was investigated, concluding that a fully hcp martensitic material has a lower wear resistance than a fully austenitic, because of the more favourable friction coefficient [10]. Co-based alloys are known to posses a high resistance to corrosion in physical media, imparted by a passive oxide film that form spontaneously on the alloy surface, and to their excellent mechanical properties. According Georgette and Davidson [11] a more stable, uniform oxide layer would be expected with a more homogenous matrix (annealed alloy) than with a highly dendritic as cast structure. Placko et al [12] found a progressive dissolution of the matrix with preferential attack of the grain boundaries and regions adjacent to carbides due to sensitization. According to Montero-Ocampo and Rodriguez [13] the low C content ASTM-F75 as cast alloys resulted in a lower release rate of corrosion products. Heat treatments may change the biocompatibility of the alloys through the modification of the electrochemical properties [14]. In the present work, three Co-26Cr-6Mo alloys with different carbon contents, produced by MIM of prealloyed powders were investigated in the as sintered and the solution annealed condition, to study the effect of the microstructure on wear and corrosion resistance.

2009-01-0020

2. Experimental Procedures A pre-alloyed Co-29Cr-6Mo (F75) powder with different carbon contents, produced by gas-atomization, was used for the production of the specimens. The powder was mixed with a proprietary binder to produce the feedstock. Test bars, according to ASTM E 8M-03 Standard Flat Unmachined Tension Test Specimen for Powder Metallurgy (P/M) Products were injection molded and debound by a two steps process (80% of the total binder content was dissolved in water, while the remaining 20% was removed by thermal decomposition). Thermal debinding was carried out in Ar/10%H2 atmosphere. The samples were sintered in vacuum furnace with holding for 1 hour in N2 backfilling at the temperatures reported in Tab. 1, and then free cooling. Tab. 1: Carbon content and sintering conditions. Carbon Content, w.t. % F75 with 0.35%C F75 with 0.23%C F75 with 0.05%C Sintering Conditions Sintering at 1300C Sintering at 1350C Sintering at 1380C

The sintered test bars were solution annealed at 1220C, 4 hours isothermal holding, in a vacuum furnace with argon backfilling and gas quenching in 8 bar nitrogen flux. The carbon content was measured by LECO CS125 after each heat treatment. Sintered density was measured by Archimedes method, resulting very close to the theoretical one in all materials (>99%). The microstructure was investigated by Light Optical Microscopy (LOM) and by Environmental Scanning Electron Microscopy (ESEM) after electrolytic etching with 94 ml distilled water, 4.5 ml HNO3 and 1.5 ml H2O2 solution for 3V and 4s. X-Ray Diffractometry (XRD) was used to investigate the cobalt matrix constitution (f.c.c./h.c.p. phases) and the type and amount of carbides. Microhardness was measured by using HV0.05, and hardness by HV30. Tensile tests were carried out in an Instron machine, at a strain rate of 0.1 s-1 and measuring strain with an axial extensometer with a gauge length of 12.5 mm. The morphology of the fracture surfaces was observed by ESEM. Wear tests: The tests were performed in a block on disc configuration using a AMSLER A135 testing machine. Co29Cr-6Mo alloy and ultrahigh molecular weight polyethylene (UHMWPE) were used as block and disc materials, respectively. Tests were carried out with a sliding speed of 0,105m/s (50 rpm) and 300N applied load, corresponding to 24MPa nominal contact stress. Water was used as lubricant. The total sliding distance was 3 km and the weight loss was measured after each test. The surface roughness of the samples was measured before and after wear tests by Hommelwerke T8000 and the results analyzed with a Turbo Roughness V 2.86 program. Corrosion tests: Two kinds of corrosion tests have been carried out, the first one was the measurement of open-circuit potential according to ISO 16429 Implants for surgery and the second one the potentiodynamic polarization. For both tests the samples were first mechanically polished using SiC grinding paper and polished with 1m diamond solution to a mirror finish and then cleaned in alcohol and dried. The OCP measurements were carried out in a thermostatically controlled cell at 37C. The electrolyte solutions employed consisted of 0.9% NaCl and as reference electrode a silver/silver chloride electrode (Ag/AgCl 3M) was used. The inert Argon gas was continuously flushed during 72 hours test. The polarization test were carried out in the same environment and temperature, but no gas was flushed in this case. The electrochemical cell was a 3 electrode cell using platinum as counter and Ag/AgCl 3M as reference. Since the open circuit potential was very instable at the beginning of immersion, all the

2009-01-0020

samples were immersed for 30 min before starting the potentiodynamic measurements. The potential range applied was from -0.25 mV vs OCP to 1 V with a scan rate of 0.1 mV/s. 2 measurements were performed for each samples and representatives curves will be reported. 3. Results and discussion

3.1. Microstructure Figures 1a and 1b show the microstructure of the as sintered and solution annealed 0.35%C alloy, respectively. The as sintered alloy contains both lamellar carbides (Cr23C6) within large eutectic cells and grain boundary carbides (Cr7C3 and M6C), as well as a fine grain boundary precipitation of sigma phase. Most of the cellular carbides of the 0.35%C alloy are effectively dissolved by solution annealing, and only a discontinuous network of grain boundary carbides (Cr23C6 and Cr7C3) is still present. The matrix of sintered material is fully h.c.p., whilst after solution annealing it contains 62% of f.c.c. which is stabilized by carbon and chromium against the transformation to martensite on cooling.

Fig. 1: Optical micrographs of as sintered (a) and solution annealed (b) 0,35%C material Figures 2a and 2b show the microstructure of the as sintered and solution annealed 0.23%C alloy, respectively. This sintered material contains again both lamellar Cr23C6 carbides and grain boundary M6C carbides and sigma phase. After solution annealing, some small spheroidized M23C6 particles (resulting form breaking-up of the lamellar carbides [2]) and a very few amount of sigma phase are still present at the grain boundary. The matrix of as sintered material is a mixture of h.c.p. and f.c.c. phases, with 57% of martensite, which increases up to 67% after solution annealing.

Fig. 2: Optical micrographs of as sintered (a) and solution annealed (b) 0,23%C material

2009-01-0020

Figures 3a and 3b show the microstructure of the as sintered and solution annealed 0.05%C alloy, respectively. The as sintered alloy comprises the metallic matrix with a discontinuous precipitation of Cr23C6 carbides on the original particles boundaries. Solution annealing is very effective on this material; only a few Cr23C6 particles are visible in the heat treated microstructure. The matrix is fully martensitic in both the as sintered and the solution annealed material.

Fig. 3: Optical micrographs of as sintered (a) and solution annealed (b) 0,05%C material 3.2. Mechanical properties The mechanical properties of all these materials, in both the as sintered and solution annealed conditions have been presented in previous papers [6, 7]; hardness and tensile properties are summarized in Table 2, and compared with the prescriptions of the ISO standard for orthopaedic implants. Table 2: Hardness and mechanical properties Tensile Test Hardness Yield Strength (HV30) U.T.S. (MPa) (MPa) 347.2 6.6 312.6 5.3 282 8 313 16 252.4 7.7 258.8 10.4 604.0 7.2 556.5 20.2 426 2 440 9 381.1 36.3 351.3 12.2 > 450 889.0 13.8 920.0 12 695 8 763 6 739.9 22.6 703.6 52.7 > 665

Samples 0.35%C As Sint. [7] 0.35%C SA [7] 0.23%C As Sint.[6] 0.23%C SA [6] 0.05%C As Sint. [7] 0.05%C SA [7] CoCrMo for implants ISO [15]

Elongation % 5.8 0.9 14.4 2.2 10 1 18 1 23.4 1.4 18.5 5.2 >8

In principle, on increasing the carbon content the strength of the alloy increases, whilst ductility decreases, because of the effect of carbides on the deformation and fracture behavior [6-8]. The effect of solution annealing depends on the carbon content. The dissolution of carbides is accomplished by changes in the matrix constitution. Since the plastic deformation of the matrix is strongly influenced by its constitution (a strain induced martensitic transformation occurs enhancing ductility [6, 7]) and by the defectiveness of martensite (which is highly faulted [16]), there isnt an unambiguous effect of solution annealing on the three materials. A detailed discussion of the mechanical properties is out from the scope of the present paper, and is proposed in previous publications [6-8]. For a global evaluation, it may be

2009-01-0020

concluded that solution annealing improves the mechanical properties of the 0.23%C and the 0.35%C materials, whilst the 0.05%C material has both better strength and ductility in the as sintered condition. With reference to the ISO standard, only the solution annealed 0.35%C material matches all the specifications, whilst the other two alloys (in particular the 0.05%C one) have a lower yield strength. 3.3. Wear resistance In figure 4 the friction coefficient tracks for the 0,35%C materials are reported, as representative of the results obtained for the other materials, too. A gradual decrease of the friction coefficient on a steady value occurs. The steady value is typical of a mixed lubricant condition, and is slightly but significantly lower for the solution annealed material.

Fig. 4: Friction coefficient x distance curve. As expected from the different hardness of the two counteracting materials, no mass loss of the Co alloys was measured, whilst some debris of UHMWPE are transferred to its surface, as figures 5 show. This phenomenon was described by Walker [9] on a cast cobalt alloy. A slight but measurable mass loss of the polymer is then induced, comparable with literature data [17]. A higher quantity of transferred polymer in the solution annealed materials is found for all materials.

Fig. 5: Wear trace for 0,05%C SA material (a) and 0,35%C sintered material (b) To quantify the wear phenomenon, the mass loss of UHMWPE and the increase in the surface roughness of the Co-alloys, which is significant of the material transfer, are considered and reported in the table 3. It has to be considered that the measure of the mass loss of the polymer, even with the use of a precision balance, is quite a hard task; therefore data must be considered as indicative.

2009-01-0020

Table 3: Mass loss of UHMWPE versus surface roughness of the Co-alloys. Samples Ra cobalt alloy (m) m UHMWPE (mg) 0.35%C As Sint. 0.028 0 0.35%C SA 0.031 -0.4 0.23%C As Sint. 0.076 -0.9 0.23%C SA 0.100 -0.6 0.05%C As Sint. 0.061 -0.4 0.05%C SA 0.076 0 The solution annealed materials cause a lower mass loss on the UHMWPE. They also evolve towards a higher surface roughness than the as sintered ones which, in combination with the lower steady value of the friction coefficient, could suggest that the transferred layer may give an additional contribution to lubrication. The 0,35%C material is the exception, because the polymer didnt present a mass loss. The data of table 3 are reported in figure 6 to compare the six materials for a global evaluation.

Fig. 6: weight loss and Ra of the six investigated materials The best wear behavior is displayed by the 0.35%C material, in the as sintered condition. It does not cause a measurable wear of the polymer and the surface roughness worsen very slightly. In presence of a lubricant, the abrasive effect of carbides is then almost completely avoided. All the other materials have a worse behavior. Among them, solution annealed 0.35%C and solution annealed 0.05%C represent two opposite situations. The former causes a measurable wear of the polymer, but roughness of the cobalt alloy does not worsen significantly; the latter does not cause a measurable mass loss of the polymer, but its surface roughness worsen significantly. This different behavior could be interpreted considering that debris of the polymer can either adhere on the metal surface (modifying roughness) or be removed by the lubricant (in this case the surface roughness does not change). What of the two options might be preferred in application is matter of discussion; in principle, the former could be preferred, which does not release debris from the tribological system and, at the same time, does not worsen the contact conditions, as the decrease of the steady friction coefficient indicates. The role of carbides, which are suspected to abrade the counterface polymer, is minimized by the lubricant. However, the comparison between as sintered and solution annealed 0.35%C seems to indicate that the localized particles after heat treatment are more abrasive than the more homogeneously distributed carbides in the as sintered material.

2009-01-0020

3.4. Corrosion tests Open circuit potential As shown in figure 7, all the materials reach a similar OCP value just after immersion that was about -0.3 and -0.2 V, but the behavior and the stabilization of the potential during time is quite different: the 0.35%C materials become much nobler during the first 10-15 hours of the immersion, showing an increase of about 0.3 and 0.5 V compared to the starting value and monitoring a change in the surface equilibrium due to the exposure environment. The 0.23%C and 0.5%C materials on the contrary do not show big changes in the potential during immersion. After this initial transient two different behaviors are visible: - a stabilization of the potential on a steady value, (0,05%C SA, 0,23%C Sint and 0,23%C SA) - a slight continuous increase of the potential, towards more noble values (0,05%C Sint., 0,35%C Sint. and 0,35%C SA).

Fig. 7: Open circuit curves for 0,35%C (a), 0,23%C (b) and 0,05%C (c) materials. The influence of the presence of carbides can be appreciated comparing the same material in the as sintered and solution annealed conditions, fig. 7, even if they have a different matrix constitution. When the carbides content increases the potential increases too. In order to consider the effect of the matrix constitution on the OCP evolution the 0,05%C SA, 0,023%C SA and 0,35%C SA materials containing respectively a decreasing amount of hcp phase and almost no carbides are compared. The difference in OCP values is quite evident and the more hcp fraction, the less noble is the electrochemical behavior. Considering samples with the same matrix phase, the comparison between 0.05%C sintered and solution annealed materials and 0,35%C sintered confirms that the higher quantity of carbides tends to enhance nobility.

2009-01-0020

Potentiodynamic curves In order to better evaluate the stability of the passivity and the electrochemical behavior of the samples potentodynamic curves were analyzed. Some representative potentiodynamic curves are reported in the figure 8.

Fig. 8: Potentiodynamic curves. All 3 materials show passive behavior in this environment and the main electrochemical parameters are summarized in table 4. Table 4: Electrochemical parameters. Samples Ecorr [V] 0.05%C SA -0.48 0.23%C SA -0.2 0.35%C As Sint. -0.24 ipass [A/cm2] 5 10-7 10-6 3 10-7 Eb [V] 0.6 0.6 0.6

The materials 0,35%C sintered and 0,23%C SA present a nobler free corrosion potential compared to 0.05%C, as found in open circuit curves. Regarding the passive corrosion density, the material with 0,23%C presents higher values compared to 0.35%C and 0.05%C but the potential of breakdown is the same for all specimens. To understand the morphology of the corrosion process, the specimens have been observed by optical microscope just after breakdown and after the evolution of corrosion process ( figures 9, 10 and 11). Based on the micrographs we can appraise: - Carbides do not oxidize; - The metal matrix oxidizes, the breakdown is due to the transpassivity and no pits are present; - The corrosive attack is localized at the carbides-matrix interface and leads to the complete removal of the carbides particles/eutectic cells. The composition of the layer is predominantly Cr2O3 oxide with some minor contribution of other oxides (Co- and Mo- oxides) [18]. According Hodgson et al. [19], the passive behavior of the CoCrMo is due to a formation the oxide layer with high content in Cr (mainly as Cr III) and smaller amount as Cr(OH)3. Moreover it may be appreciated that the corrosion of the 0.05%C occurs by localized attacks, while the 0.23%C samples show an intergranular corrosion morphology. The 0.35%C shows a large corroded area corresponding carbide/eutectic phase. Therefore, although more carbides nobler the electrochemical behavior, the presence of the carbide/eutectic phase lead to a different corrosion morphology when the traspassivity state is reached.

2009-01-0020

Fig. 9: Corrosion micrographs of 0,35%C after breakdown (a) and the evolution of corrosion process (b)

Fig. 10: Corrosion micrographs of 0,23%C after breakdown (a) and the evolution of corrosion process (b)

Fig. 11: Corrosion micrographs of 0,05%C after breakdown (a) and the evolution of corrosion process (b) 4. Conclusions In this work the influence of microstructure on the corrosion and wear properties of a Co-Cr-Mo alloy produced by Metal Injection Molding, was investigated. The materials have been sintered to full density and heat treated to dissolve carbides. The matrix constitution changed correspondingly, and hcp martensite reverts partially to fcc austenite.

2009-01-0020

Wear tests were carried out against a UHMWPE under lubricated conditions. The wear mechanism is characterized by the transfer of wear debris from the polymer to the metallic alloy, and is quantified by the mass loss of the former and the increase in the roughness of the latter. The material with the best wear properties is as sintered 0,35%C alloy, which does not cause a measurable mass loss of the polymer and display only a very slight increase in the surface roughness. Corrosion tests were carried out by measuring the Open Circuit Potential. All the materials have a fast transition to a noble behavior, with an effect of both the matrix constitution and the carbide amount. In particular, the more hcp fraction, the less noble is the electrochemical behavior and a higher quantity of carbides tends to enhance nobility. On forcing corrosion by potentiodynamic tests, carbides tend to localize the attack responsible for transpassivation. For a global evaluation of the six materials investigated, wear and corrosion would suggest the use of the as sintered 0.35%C material. However, its mechanical properties do not match the ISO specifications, since ductility is strongly depressed by carbides. After solution annealing ductility increases above the minimum value required. In this condition, corrosion behavior remains quite good, whilst the residual isolated carbides tend to enhance the abrasion of the polymer. Since the mass loss of the polymer is very low, and compatible with the application, this material could represent the best solution for the production of the orthopedic prostheses by MIM of prealloyed powders. 5. References 1. T. Kilner, R.M. Pilliar, G.C. Weatherly, and C. Allibert, J. Biomedical Mater. Res., 16(1982)63. 2. J.W. Weeton and R.A. Signorelli, Technical Note 3109, National Advisory Committee for Aeronautics, 1954. 3. F.J. Shortsleeve and M.E. Nicholson, Transactions, American Society for Metals 43(1951)142. 4. H. S. Dobbs, J. L. M. Robertson, Journal of Materials Science 18(1983)391 5. H. Mancha et al., Journal of Materials Synthesis and Processing 4(4)(1996)217 6. P.V. Muterlle, M. Zendron, M. Perina, R. Bardini, A. Molinari. Effect of sintering temperature and heat treatment on microstructure and tensile properties of Co-Cr-Mo alloy with 0.23%C produced by MIM (2009) in press on Journal of Materials Science. 7. Muterlle P. V., DIncau M., Perina M., Bardini R., Molinari A. Advances in Powder Metallurgy & Particulate Materials 4(2008)183. 8. P.V. Muterlle, M. Perina, M. Mantovani, A. Molinari, Metal Powder Report, 64(2)(2009)30. 9. P.S. Walker, G.W. Blunn and P.A. Lilley, Journal of Biomedical Materials Research 33(1996)159. 10. A. J. Saldivar Garcia, H. F. Lopez, Journal of Biomedical Materials Research 74A(2005)269 11. F.S. Georgette, J.A. Davidson, Journal of Biomedical Materials Research 20(1986)1229. 12. H.E. Placko, S.A. Brown, J.H. Payer, Journal of Biomedical Materials Research 39(1998)292. 13. C. Montero-Ocampo and A.S. Rodriguez, Journal of Biomedical Materials Research 29(1995)441 14. C.V. Vidal and A.I. Muoz, Electrochimica Acta 54(2009)1798. 15. ISO 5832-4 Implants for surgery - Metallic materials - Cobalt-Crhomium-Molybdenum cating alloys, second edition (1996). 16. K. Rajan, Metallurgical Transactions 13A(1982)1161 17. Dr. S. Mischler, Tribology and Implants. Cours Biomatriaux, novembre 2006. 18. I. Milosev, H. Strehblow, Electrochimica Acta 48(2003)2767. 19. A.W.E. Hodgso et al., Electrochimica Acta 49(2004)2167.

También podría gustarte