Next Article in Journal
Effects of Thermomechanical Treatment on Phase Transformation of the Ti50Ni49W1 Shape Memory Alloy
Previous Article in Journal
Detailed Thermo-Kinematic Analysis of Face Grinding Operations with Straight Wheels
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Evaluation of Biomedical Ti/ZrO2 Joint Brazed with Pure Au Filler: Microstructure and Mechanical Properties

1
State Key Laboratory of Advanced Welding and Joining, Harbin Institute of Technology, Harbin 150001, China
2
Shandong Provincial Key Lab of Special Welding Technology, Harbin Institute of Technology at Weihai, Weihai 264209, China
3
State Key Laboratory of New Brazing Materials and Technology, Zhengzhou Research Institute of Mechanical Engineering Co., Ltd., Zhengzhou 450001, China
4
Shandong Institute of Shipbuilding Technology, Weihai 264209, China
*
Author to whom correspondence should be addressed.
Metals 2020, 10(4), 526; https://doi.org/10.3390/met10040526
Submission received: 19 March 2020 / Revised: 11 April 2020 / Accepted: 16 April 2020 / Published: 18 April 2020

Abstract

:
Titanium and zirconia (ZrO2) ceramics are widely used in biomedical fields. This study aims to achieve reliable brazed joints of titanium/ZrO2 using biocompatible Au filler for implantable medical products. The effects of brazing temperature and holding time on the interfacial microstructures and mechanical properties of titanium/Au/ZrO2 joints were fully investigated by scanning electron microscopy (SEM), energy-dispersive spectrometry (EDS) and X-ray diffraction (XRD). The results indicated that the typical interfacial microstructure of the titanium/Au/ZrO2 joint was titanium/Ti3Au layer/TiAu layer/TiAu2 layer/TiAu4 layer/TiO layer/ZrO2 ceramic. With an increasing brazing temperature or holding time, the thickness of the Ti3Au + TiAu + TiAu2 layer increased gradually. The growth of the TiO layer was observed, which promoted metallurgical bonding between the filler metal and ZrO2 ceramic. The optimal shear strength of ~35.0 MPa was obtained at 1150 °C for 10 min. SEM characterization revealed that cracks initiated and propagated along the interface of TiAu2 and TiAu4 reaction layers.

1. Introduction

As one of the most popular biomedical metallic materials, titanium and titanium alloys have been widely used to produce bone trauma products, artificial joints, cardiovascular stents, dental implants and other medical products, owing to their low density, low elastic modulus, non-toxic behavior, good corrosion resistance and excellent biocompatibility [1,2,3,4,5,6,7,8]. In recent decades, the production and application of zirconia bioceramics have developed rapidly. The properties that it processes, such as its high hardness, high wear resistance, excellent biocompatibility and aesthetic effect make it suitable for surgical implant fabrication, especially for implants in the field of prosthodontics [9,10,11,12,13]. In many applications, metal–ceramic hybrid components are desired for the manufacturing of implantable medical products—for example, dental implants, micro-stimulators and so on [14,15,16].
At present, the main joining methods for metal–ceramic components in implant manufacturing are cementation and mechanical bonding [17,18,19,20]. However, the joints of metal–ceramic composites constructed via these bonding methods display a low strength and are easy to loosen in practice, which can shorten the life of implants and lead to potential dangers in applications. Brazing, a bonding technology with the advantages of convenience, cost-effectiveness and high quality, has been widely employed for joining metals and ceramics [21,22,23]. Sharma et al. [24,25] realized the brazing of Ti-6Al-4V to ZrO2 successfully using Ag-Cu-In-Ti active filler and Ag-Cu-Ti composite fillers. ZrO2 was boned under the influence of active Ti from filler, which was absorbed in the surface pores through capillary action at the ZrO2 surface. Fu et al. [26] and Bian et al. [27,28] used Sn-Ti and SnAgCu-Ti active fillers to achieve the bonding of zirconia ceramics. The TiOx compounds were observed on the ZrO2 side. Smorygo et al. [16] used Cu-Ag-Ti filler to achieve the bonding of titanium to a zirconia ceramic by forming a layer of TiOx with a thickness of 3–4 μm between the filler and ZrO2 sample. Moreover, the authors pointed out that the brazing temperature and holding time had great influences on the evolution of the bond layer microstructure and the fracture behavior. Feng et al. [29] and Dai et al. [30] utilized AgCu and AgCu composite fillers to achieve the bonding of a titanium alloy to a zirconia ceramic. The Ti from the substrates crossed the brazing seam, accumulated on the ZrO2 and reacted with the ZrO2 to form a TiO layer. In other words, the key to obtaining reliable bonding is the formation of Ti-O compounds adjacent to the ceramics. There are two main ways for Ti to form Ti-O reported in the literature, namely by adding Ti into the filler metal and by diffusing it from the base metal. The issue is that the fillers that are usually used to braze metal and ceramics mainly contain toxic elements, making them unsuitable for use in the bonding of biomedical joints. One potential solution is to use Au, a biocompatible element, as the filler metal [31,32,33,34,35,36]. In addition, the melting point of Au is much lower than that of titanium and zirconia ceramics, and its ductility is high, which is conducive to the stress relief of metal–ceramic brazed joints. Furthermore, according to the binary alloy phase diagram of Ti-Au [37], Ti can react with Au. Thus, when adopting pure Au as the filler to bond titanium or titanium alloys to ceramics with appropriate brazing parameters, Ti can cross the brazing seam to react with ceramics and realize the bonding of titanium or titanium to ceramics. Bian et al. [38] adopted Au foil to braze titanium and alumina ceramic and a good brazed joint was successfully obtained for implantable devices, where the formation of a Ti–O layer adjacent to alumina was deduced but not definitely identified. Up to now, studies on brazing metals and ceramics for biomedical applications are still rare.
In this study, the reliable brazing of biomedical titanium to zirconia ceramic was achieved by adopting biocompatible Au filler. The typical microstructure of the brazed joints was analyzed. Detailed investigations into the effects of brazing temperature and holding time on the microstructural evolution, Ti–O compound layer and mechanical properties were conducted. The shear strength of joints was tested, and fracture analyses were conducted to understand the mechanisms of the fractures.

2. Experimental Materials and Methods

Commercial pure titanium of purity 99.6 wt.%, mainly doped with 0.2 wt.% Fe and 0.18 wt.% O, provided by Kunshan Bitaita Metal Products Co., Ltd., Kunshan, China, was cut into 20 mm × 10 mm × 2 mm pieces. Figure 1a,b shows the microstructure and XRD pattern of pure titanium (according to PDF#00-044-1294), respectively. It was clearly seen that the pure titanium mainly consisted of α-Ti with an equiaxed structure. Sintered 3 mol% yttria-stabilized zirconia, supplied by Shanghai Unite Technology Co., Ltd., Shanghai, China, was cut into 5 mm cubes using a diamond cutter. The back-scattered electron (BSE) image of ZrO2 was shown in Figure 1c. Au foil with a purity of 99.99% and thickness of 50 µm, which was used in the experiment, was supplied by KYKY Technology Co., Ltd., Beijing, China.
Prior to vacuum brazing, the surface of the titanium to be brazed was ground to a grit of 3000 mesh by SiC grinding paper. Both substrates and the Au foil were cleaned using an ultrasonic bath in acetone for 15 min, followed by air blowing. Then, the Au foils were sandwiched between the substrates, as shown in Figure 1d. Brazing was performed in a vacuum furnace with a vacuum of 1.3 × 10−3 Pa. The assembly was firstly heated to 1000 °C for 10 min from an ambient temperature, at a heating rate of 20 °C/min, and then two groups of experiments were designed in order to investigate the effects of brazing temperature and holding time on the microstructures and mechanical properties of the brazed joints: in one experiment, the temperature continued to increase to the brazing temperature (1110–1190 °C) at a rate of 10 °C/min, with the holding time fixed at 10 min; in the other set of experiments, the temperature continued to increase to 1150 °C at a rate of 10 °C/min, with the holding time varying from 5 to 30 min. Subsequently, the specimens were cooled down to 300 °C at a rate of 5 °C/min. Finally, the assembly was spontaneously cooled to room temperature in the furnace.
After the experiments, the cross-sections of brazed joints, which were obtained by cutting the specimens perpendicular to the brazed interface using a diamond saw, were polished for microstructural observations, and they were characterized with an SEM (MERLIN Compact, ZEISS, Stuttgart, Germany) in BSE mode, equipped with an EDS (Octane Plus, EDAX, Mahwah, NJ, USA) to analyze the composition of various reaction phases. The shear tests were performed with a universal testing machine (Instron 5967, Instron, Boston, MA, USA) at a constant rate of 1 mm/min at room temperature. The experimental data were averaged from at least five specimens after removing the outliers for each parameter. After the shear test, three randomly selected fractured specimens were analyzed by SEM in BSE mode and XRD (DX-2700, Dandong Haoyuan Instrument Co., Ltd., Dandong, China), equipped with Cu-Kα (λ = 0.154 nm), at operating parameters of 40 KV and 30 mA to identify the fracture path.

3. Results and Discussion

3.1. Typical Interfacial Microstructure of Titanium/Au/ZrO2 Joint

Figure 2 showed the typical microstructure and the main element distribution of the titanium/Au/ZrO2 joint brazed at 1150 °C for 10 min. As shown in Figure 2a–b, a sound joint without any microcracks or pores was obtained, and the joint could be divided into five zones based on its different microscopic morphologies. According to the elemental distribution along the red line shown in Figure 2c, it can be seen that the content of Ti decreased gradually from the titanium substrate to the ZrO2 ceramic, while an opposite trend for Au was observed. It was worth noting that there were four platforms in the variation curve of both Ti and Au elements, which corresponded to Zones I, II, III and IV in Figure 2b. Combined with the elemental distribution maps in Figure 2d–g, it was deduced that Ti dissolved and diffused to the molten Au. At the same time, Au also diffused to the Ti substrate. From the phase diagram of Au-Ti [37], it can be seen that Ti and Au form TixAuy intermetallic compounds (IMCs) easily. In addition, the diffusion of Ti across the brazing seam occurred, with Ti segregation on the ZrO2 ceramic side also observed, forming Ti-O compounds via the following reaction: Ti+ ZrO2 → TiOx + ZrO2−x [39,40,41]. Eventually, Zone V was formed via metallurgical bonding between the filler metal and the ceramic.
In order to reveal more details of each zone in the titanium/Au/ZrO2 joint, the highly magnified interfacial microstructures of Zones I–V are shown in Figure 3. The EDS chemical compositions of each spot in Figure 3 are listed in Table 1. The EDS analyses of Zones I–IV showed that these zones mainly contained Ti and Au. According to the molar ratio of Ti/Au and the Ti-Au binary phase diagram [37], it can be concluded that the zones from I to VI were Ti3Au phase (Spot A), TiAu phase (Spot B), TiAu2 phase (Spot C) and TiAu4 phase (Spot D), respectively. When the brazing temperature exceeded the melting point of Au (1064 °C), the Au foil began to melt and the interdiffusion of Ti and Au occurred, both of which were driven by the concentration gradient. Then, Ti reacted with Au to form Ti3Au, TiAu, TiAu2 and TiAu4 IMCs due to the decreasing concentration gradient of Ti [38]. Through the cooling process, the Ti3Au, TiAu, TiAu2 and TiAu4 layers formed. The thicknesses of these layers, on average, were ~4.8 μm, 5.3 μm, 14.2 μm and 10.2 μm combined with Figure 2a–b. The chemical composition of Spot E in Zone V detected by EDS analysis showed that the titanium/oxygen atomic ratio was about 1:1. Combined with the results of Ti reacting with ZrO2 ceramics in previous studies [42,43], it can be deduced that the black layer (Zone V) next to the ZrO2 side was a TiO layer. Therefore, the active Ti from the substrate had two effects during brazing: one was to react with Au to form Ti–Au compounds, and the other was to react with the ZrO2 ceramic to form a metallurgical bond.
Based on the above analyses, it can be concluded that the representative brazing microstructure of the titanium/Au/ZrO2 joint brazed at 1150 °C for 10 min was titanium/Ti3Au layer/TiAu layer /TiAu2 layer /TiAu4 layer /TiO layer /ZrO2 ceramic.

3.2. Effects of Brazing Parameters on the Interfacial Microstructure of the Titanium/Au/ZrO2 Joints

It is acknowledged that the brazing temperature plays an important role in the evolution of interfacial microstructure [16,29]. The microstructural evolutions of brazed joints at 1110–1190 °C at intervals of 20 °C are shown in Figure 4a–e. It can be seen that the zones and phases in the joints are consistent with the typical interfacial microstructure. However, with a rise in the brazing temperature, the thickness of Ti3Au+TiAu+TiAu2 layers increased gradually from 19.2 μm to 24.3μm, owing to the enhanced diffusion of Ti and Au with the increasing temperature, as shown in Figure 4a–c. With the brazing temperature increasing further, as shown in Figure 4d–e, the thickness of Ti3Au + TiAu + TiAu2 layers showed no obvious change due to the Ti–Au reaction layers hindering the further diffusion of Ti and Au by acting as barriers.
In order to further analyze the microstructural evolutions of the TiO layer in the titanium/Au/ZrO2 joint, the highly magnified microstructures of Zone V are shown in Figure 5. When the brazing temperature was low (e.g., 1110 °C), only a limited number of Ti atoms diffused to the ZrO2 surface and reacted with O atoms from ZrO2 to form a thin layer of TiO, which was not obviously observed by SEM, as shown in Figure 5a. With the increase in temperature, more sufficient Ti atoms diffused to and accumulated on the ZrO2 surface. Thus, the thickness of the TiO layers increased gradually, as shown in Figure 5b–e.
It was well known that holding time is the other important factor affecting brazing quality, besides brazing temperature. Figure 6a–e showed the BSE images of the interfacial microstructure of the titanium/Au/ZrO2 joint brazed at 1150 °C for 5–30 min, respectively. Notably, the joints still consisted of five zones and the phases were consistent with the typical interfacial microstructure shown in Figure 2. With a longer holding time, the diffusion of Ti atoms was more sufficient and the thickness of Ti3Au + TiAu + TiAu2 layers increased gradually from 22.6 μm to 30.3 μm. Additionally, the TiO layer showed no significant changes, except for a slight increase in thickness.
In summary, the effects of the brazing parameters on the microstructural evolution of the joints can be summarized as follows: when the brazing temperature exceeded the melting point of the Au foil, the Au foil started to melt and spread on the surface of the titanium substrate. Meanwhile, Ti dissolved and diffused to the molten Au which was driven by the concentration gradient of Ti. On the one hand, Ti reacted with Au to form Ti–Au IMCs. In the cooling process, Ti3Au, TiAu, TiAu2 and TiAu4 layers formed simultaneously in the brazing seam due to the decreasing concentration gradient of Ti. With the temperature increasing, the diffusion of Ti and Au was facilitated, and the concentration gradient of Ti grew. Therefore, the thickness of the Ti3Au + TiAu + TiAu2 layer increased gradually. However, with the brazing temperature increasing further, Ti–Au IMC layers prevented the further interdiffusion of Ti and Au atoms, which resulted in an indiscernible change in the thickness of Ti3Au + TiAu + TiAu2 layers. Similar to the effect of the brazing temperature, with a longer holding time, the diffusion of Ti and Au atoms was more sufficient, and the thickness of the Ti3Au + TiAu + TiAu2 layer increased. On the other hand, redundant Ti diffused to the ZrO2 ceramic side crossing the brazing seam and accumulated on the surface of the ZrO2 ceramic. Active Ti could partially capture oxygen from ZrO2 to form a TiO compound. When the brazing temperature was low, the diffusion speed of the Ti atoms was slow and the number of Ti atoms that diffused to the ZrO2 surface was limited. Therefore, the TiO layer was too thin to be observed by SEM. With an increase in temperature, a more sufficient diffusion of Ti atoms to the ZrO2 surface occurred, which then accumulated on the ZrO2 surface and reacted with the O atoms from ZrO2. Therefore, the thickness of the TiO layers increased gradually. However, with a longer holding time, the Ti-Au IMC layers hindered the diffusion of more Ti atoms to ZrO2. Thus, the TiO layer displayed no significant change, except for a slight increase in thickness.

3.3. Mechanical Properties and Fracture Morphology of Titanium/Au/ZrO2 Joint

To evaluate the mechanical properties of the brazed joints, shear tests at room temperature were carried out. Figure 7a,b illustrated the effect of the brazing temperature and holding time on the average shear strength of titanium/Au/ZrO2 joints, respectively. It can be seen clearly that both temperature and holding time influenced the shear strength significantly. The shear strength improved with the increase in brazing temperature and holding time, until the optimal shear strength of ~35.0 MPa was reached at 1150 °C for 10 min, and then decreased.
In order to further investigate the fracture mechanisms of the joints after the shear test, the fracture paths and their magnified fractographs at different brazing temperature and holding times are shown in Figure 8(a1–a2,b1–b2,c1–c2). Moreover, the corresponding XRD spectra (ZrO2: PDF#01-088-1007, TiAu2: PDF#00-029-0651 and TiAu4: PDF#04-004-9165) on the ZrO2 side are illustrated in Figure 8(a3,b3,c3). When the joints were brazed at a low temperature or for a short holding time—for example, 1110 °C/10 min—the diffusion of Ti and Au atoms was slow and insufficient. Thus, only a limited number of Ti atoms diffused to the ZrO2 surface and reacted with ZrO2 to form a TiO layer. As a result, the TiO layer was too thin, resulting in weak metallurgical bonding between ZrO2 and the brazing alloy. As a result, cracks initiated in the ZrO2/TiAu4 interface and propagated to the brazing seam, mainly causing cracks on the TiAu4/TiAu2 (Zone IV/III) interface, as shown in Figure 8(a1–a2). The corresponding XRD spectrum on the ZrO2 side in Figure 8(a3) mainly contained TiAu4 and TiAu2 phases, which further proves the above analysis. With the temperature increasing, the diffusion of Ti and Au atoms accelerated and more Ti atoms accumulated on the ZrO2 surface, forming a thicker and more continuous TiO layer in order to achieve better metallurgical bonding on the ZrO2/brazing seam interface. In this case, the shear strength increased gradually. Figure 8(b1–b2) shows the crack path and fractography obtained at 1150 °C for 10 min. The thickness of the TiO layer was the most moderate and the joint achieved the maximum shear strength. The joints mainly cracked on the brazing seam. Combined with the XRD spectrum in Figure 8(b3), it can be concluded that the fracture path was located on the TiAu4/TiAu2 interface. With the brazing temperature and holding time further increasing, the TiO layer became thicker and the residual stress between ZrO2 and the brazing seam increased. As a result, the joints mainly cracked on the brazing seam and the ZrO2/TiAu4 interface, as shown in Figure 8(c1–c3) for the joint brazed at 1150 °C for 30min. Based on the above analysis, it can be deduced that the formation and proper thickness of the TiO layer was crucial to the shear strength of the joints.

4. Conclusions

In this study, the reliable bonding of biomedical titanium/ZrO2 was successfully achieved using Au foil. The typical interfacial microstructure of the brazed joint was characterized. The effects of brazing temperature and holding time on the interfacial microstructure and mechanical properties of the joints were investigated in detail. In conclusion:
(1) Ti3Au, TiAu, TiAu2, TiAu4 and TiO phases were formed in the brazed joint. The typical interfacial microstructure of the titanium/Au/ZrO2 joint was titanium/Ti3Au layer/TiAu layer/TiAu2 layer/TiAu4 layer/TiO layer/ZrO2 ceramic. The thicknesses of these layers, on average, were ~4.8 μm, 5.3 μm, 14.2 μm and 10.2 μm;
(2) The brazing temperature and holding time had significant effects on the interfacial microstructure and mechanical properties of the brazed joints. With a higher brazing temperature or a longer holding time, the diffusion of Ti and Au was accelerated, and the thickness of the Ti3Au + TiAu + TiAu2 layers increased gradually. The TiO layer thickened gradually and promoted metallurgical bonding between the brazing alloy and the ZrO2 ceramic;
(3) The joint brazed at 1150 °C for 10 min had an optimal shear strength of ~35.0 MPa, and a TiO layer with a modest thickness. A crack was initiated and propagated along the interface of the TiAu2 and TiAu4 reaction layers.

Author Contributions

Conceptualization, X.S. and W.L.; formal analysis, Y.L.; funding acquisition, H.B. and X.S.; investigation, Y.L., W.F. and H.N.; project administration, J.F.; resources, H.N.; supervision, H.B. and J.F.; validation, W.F.; visualization, W.F.; writing—original draft, Y.L.; writing—review & editing, all authors. All authors have read and agreed to the published version of the manuscript.

Funding

This project is supported by the National Natural Science Foundation of China (Grant Nos. 51905127 and 51775138) and the Natural Science Foundation of Shandong Province (Nos. ZR2019PEE042).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Xu, L.; Li, J.; Xu, X.; Lei, X.; Zhang, K.; Wu, C.; Zhang, Z.; Shi, X.; Wang, X.; Ding, J. A Novel Cytocompatibility Strengthening Strategy of Ultrafine-Grained Pure Titanium. ACS Appl. Mater. Interfaces 2019, 11, 47680–47694. [Google Scholar] [CrossRef]
  2. Kumar, D.D.; Kaliaraj, G.S. Multifunctional zirconium nitride/copper multilayer coatings on medical grade 316L SS and titanium substrates for biomedical applications. J. Mech. Behav. Biomed. Mater. 2018, 77, 106–115. [Google Scholar] [CrossRef]
  3. Gao, C.; Li, C.; Wang, C.; Qin, Y.; Wang, Z.; Yang, F.; Liu, H.; Chang, F.; Wang, J. Advances in the induction of osteogenesis by zinc surface modification based on titanium alloy substrates for medical implants. J. Alloys Compd. 2017, 726, 1072–1084. [Google Scholar] [CrossRef]
  4. Kaur, M.; Singh, K. Review on titanium and titanium based alloys as biomaterials for orthopaedic applications. Mater. Sci. Eng. C Mater. Biol. Appl. 2019, 102, 844–862. [Google Scholar] [CrossRef] [PubMed]
  5. Mishnaevsky, L.; Levashov, E.; Valiev, R.Z.; Segurado, J.; Sabirov, I.; Enikeev, N.; Prokoshkin, S.; Solov’yov, A.V.; Korotitskiy, A.; Gutmanas, E.; et al. Nanostructured titanium-based materials for medical implants: Modeling and development. Mater. Sci. Eng. R Rep. 2014, 81, 1–19. [Google Scholar] [CrossRef]
  6. Korobkova, A.; Kazakbiev, A.; Zhukova, Y.; Sheremetyev, V.; Dubinskiy, S.; Filonov, M. Surface treatment of bulk and porous materials based on superelastic titanium alloys for medical implants. Mater. Today Proc. 2017, 4, 4664–4669. [Google Scholar] [CrossRef]
  7. Liu, L.; Xu, J.; Jiang, S. Nanocrystalline β-Ta Coating Enhances the Longevity and Bioactivity of Medical Titanium Alloys. Metals 2016, 6, 221. [Google Scholar] [CrossRef] [Green Version]
  8. Reck, A.; Zeuner, A.T.; Zimmermann, M. Fatigue Behavior of Non-Optimized Laser-Cut Medical Grade Ti-6Al-4V-ELI Sheets and the Effects of Mechanical Post-Processing. Metals 2019, 9, 843. [Google Scholar] [CrossRef] [Green Version]
  9. Ayoub, G.; Veljovic, D.; Zebic, M.L.; Miletic, V.; Palcevskis, E.; Petrovic, R.; Janackovic, D. Composite nanostructured hydroxyapatite/yttrium stabilized zirconia dental inserts–The processing and application as dentin substitutes. Ceram. Int. 2018, 44, 18200–18208. [Google Scholar] [CrossRef]
  10. Marques, A.; Miranda, G.; Faria, D.; Pinto, P.; Silva, F.; Carvalho, O. Novel design of low modulus high strength zirconia scaffolds for biomedical applications. J. Mech. Behav. Biomed. Mater. 2019, 97, 375–384. [Google Scholar] [CrossRef]
  11. Dandoulaki, C.; Rigos, A.E.; Kontonasaki, E.; Karagiannis, V.; Kokoti, M.; Theodorou, G.S.; Papadopoulou, L.; Koidis, P. In vitro evaluation of the shear bond strength and bioactivity of a bioceramic cement for bonding monolithic zirconia. J. Prosthet. Dent. 2019, 122, e1–e167. [Google Scholar] [CrossRef] [PubMed]
  12. Stefanic, M.; Kosmač, T. β-TCP coatings on zirconia bioceramics: The importance of heating temperature on the bond strength and the substrate/coating interface. J. Eur. Ceram. Soc. 2018, 38, 5264–5269. [Google Scholar] [CrossRef]
  13. Volceanov, E.; Popa, C.G.; Volceanov, A.; Ciuca, S. Chemical stability in artificial saliva of zirconia bioceramics. Rev. Romana Mater. Rom. J. Mater. 2019, 49, 313–321. [Google Scholar]
  14. Cantner, F.; Cacaci, C.; Mucke, T.; Randelzhofer, P.; Hajto, J.; Beuer, F. Clinical performance of tooth- or implant-supported veneered zirconia single crowns: 42-month results. Clin. Oral Investig. 2019, 23, 4301–4309. [Google Scholar] [CrossRef]
  15. Parchanska-Kowalik, M.; Wolowiec-Korecka, E.; Klimek, L. Effect of chemical surface treatment of titanium on its bond with dental ceramics. J. Prosthet. Dent. 2018, 120, 470–475. [Google Scholar] [CrossRef]
  16. Smorygo, O.; Kim, J.S.; Kim, M.D.; Eom, T.G. Evolution of the interlayer microstructure and the fracture modes of the zirconia/Cu–Ag–Ti filler/Ti active brazing joints. Mater. Lett. 2007, 61, 613–616. [Google Scholar] [CrossRef]
  17. Yadav, J.S.; Dabas, N.; Bhargava, A.; Malhotra, P.; Yadav, B.; Sehgal, M. Comparing two intraoral porcelain repair systems for shear bond strength in repaired cohesive and adhesive fractures, for porcelain-fused-to-metal restorations: An in vitro study. J. Indian Prosthodont. Soc. 2019, 19, 362–368. [Google Scholar] [CrossRef]
  18. Haselton, D.R.; Diaz-Arnold, A.M.; Dunne, J.T., Jr. Shear bond strengths of 2 intraoral porcelain repair systems to porcelain or metal substrates. J. Prosthet. Dentistry 2001, 86, 526–531. [Google Scholar] [CrossRef]
  19. Han, X.; Sawada, T.; Schille, C.; Schweizer, E.; Scheideler, L.; Geis-Gerstorfer, J.; Rupp, F.; Spintzyk, S. Comparative Analysis of Mechanical Properties and Metal-Ceramic Bond Strength of Co-Cr Dental Alloy Fabricated by Different Manufacturing Processes. Materials 2018, 11, 1801. [Google Scholar] [CrossRef] [Green Version]
  20. Zhou, Y.; Wei, W.; Yan, J.; Liu, W.; Li, N.; Li, H.; Xu, S. Microstructures and metal-ceramic bond properties of Co-Cr biomedical alloys fabricated by selective laser melting and casting. Mater. Sci. Eng. A 2019, 759, 594–602. [Google Scholar] [CrossRef]
  21. Li, C.; Huang, C.; Chen, L.; Si, X.; Chen, Z.; Qi, J.; Huang, Y.; Feng, J.; Cao, J. Microstructure and mechanical properties of the SiC/Nb joint brazed using AgCuTi+B4C composite filler metal. Int. J. Refract. Met. Hard Mater. 2019, 85, 105049. [Google Scholar] [CrossRef]
  22. Way, M.; Willingham, J.; Goodall, R. Brazing filler metals. Int. Mater. Rev. 2020, 65, 257–285. [Google Scholar] [CrossRef]
  23. Hu, S.P.; Hu, T.Y.; Lei, Y.Z.; Song, X.G.; Liu, D.; Cao, J.; Tang, D.Y. Microstructural evolution and mechanical properties of vacuum brazed Ti2AlNb alloy and Ti60 alloy with Cu75Pt filler metal. Vacuum 2018, 152, 340–346. [Google Scholar] [CrossRef]
  24. Sharma, A.; Kee, S.H.; Jung, F.; Heo, Y.; Jung, J.P. Compressive Strength Evaluation in Brazed ZrO2/Ti6Al4V Joints Using Finite Element Analysis. J. Mater. Eng. Perform. 2016, 25, 1722–1728. [Google Scholar] [CrossRef]
  25. Sharma, A.; Ahn, B. Brazeability, Microstructure, and Joint Characteristics of ZrO2/Ti-6Al-4V Brazed by Ag-Cu-Ti Filler Reinforced with Cerium Oxide Nanoparticles. Adv. Mater. Sci. Eng. 2019, 2019, 1–11. [Google Scholar] [CrossRef] [Green Version]
  26. Fu, W.; Song, X.; Passerone, A.; Hu, S.; Bian, H.; Zhao, Y.; Wang, M.; Valenza, F. Interactions, joining and microstructure of Sn-Ti/ZrO2 system. J. Eur. Ceram. Soc. 2019, 39, 1525–1531. [Google Scholar] [CrossRef]
  27. Bian, H.; Fu, W.; Lei, Y.Z.; Song, X.G.; Liu, D.; Cao, J.; Feng, J.C. Wetting and low temperature bonding of zirconia metallized with Sn0.3Ag0.7Cu-Ti alloys. Ceram. Int. 2018, 44, 11456–11465. [Google Scholar] [CrossRef]
  28. Bian, H.; Zhou, Y.; Song, X.; Hu, S.; Shi, B.; Kang, J.; Feng, J. Reactive wetting and interfacial characterization of ZrO2 by SnAgCu-Ti alloy. Ceram. Int. 2019, 45, 6730–6737. [Google Scholar] [CrossRef]
  29. Feng, J.; Dai, X.; Wang, D.; Li, R.; Cao, J. Microstructure evolution and mechanical properties of ZrO2/TiAl joints vacuum brazed by Ag–Cu filler metal. Mater. Sci. Eng. A 2015, 639, 739–746. [Google Scholar] [CrossRef]
  30. Dai, X.; Cao, J.; Liu, J.; Wang, D.; Feng, J. Interfacial reaction behavior and mechanical characterization of ZrO2/TC4 joint brazed by Ag–Cu filler metal. Mater. Sci. Eng. A 2015, 646, 182–189. [Google Scholar] [CrossRef]
  31. Jain, P.K. Gold Nanoparticles for Physics, Chemistry, and Biology. Edited by Catherine Louis and Olivier Pluchery. Angew. Chem. Int. Ed. 2014, 53, 1197. [Google Scholar] [CrossRef]
  32. Manickam, P.; Vashist, A.; Madhu, S.; Sadasivam, M.; Sakthivel, A.; Kaushik, A.; Nair, M. Gold nanocubes embedded biocompatible hybrid hydrogels for electrochemical detection of H2O2. Bioelectrochemistry 2020, 131, 107373. [Google Scholar] [CrossRef] [PubMed]
  33. Iram, F.; Iqbal, M.S.; Khan, I.U.; Rasheed, R.; Khalid, A.; Khalid, M.; Aftab, S.; Shakoori, A.R. Synthesis and Biodistribution Study of Biocompatible (198)Au Nanoparticles by use of Arabinoxylan as Reducing and Stabilizing Agent. Biol. Trace Elem. Res. 2020, 193, 282–293. [Google Scholar] [CrossRef] [PubMed]
  34. Shuai, H.L.; Wu, X.; Huang, K.J.; Zhai, Z.B. Ultrasensitive electrochemical biosensing platform based on spherical silicon dioxide/molybdenum selenide nanohybrids and triggered Hybridization Chain Reaction. Biosens. Bioelectron. 2017, 94, 616–625. [Google Scholar] [CrossRef]
  35. Manohar, N.; Reynoso, F.J.; Diagaradjane, P.; Krishnan, S.; Cho, S.H. Quantitative imaging of gold nanoparticle distribution in a tumor-bearing mouse using benchtop x-ray fluorescence computed tomography. Sci. Rep. 2016, 6, 22079. [Google Scholar] [CrossRef] [Green Version]
  36. Chen, Z.; Meng, H.; Xing, G.; Chen, C.; Zhao, Y.; Jia, G.; Wang, T.; Yuan, H.; Ye, C.; Zhao, F.; et al. Acute toxicological effects of copper nanoparticles in vivo. Toxicol. Lett. 2006, 163, 109–120. [Google Scholar] [CrossRef]
  37. Murray, J.L. The Au-Ti (Gold-Titanium) system. Bull. Alloy. Phase Diagr. 1983, 4, 278–283. [Google Scholar] [CrossRef]
  38. Bian, H.; Song, X.; Hu, S.; Lei, Y.; Jiao, Y.; Duan, S.; Feng, J.; Long, W. Microstructure Evolution and Mechanical Properties of Titanium/Alumina Brazed Joints for Medical Implants. Metals 2019, 9, 644. [Google Scholar] [CrossRef] [Green Version]
  39. Lin, K.F.; Lin, C.C. Interfacial reactions between zirconia and titanium. Scr. Mater. 1998, 39, 1333–1338. [Google Scholar] [CrossRef]
  40. Hanson, W.B.; Ironside, K.I.; Fernie, J.A. Active metal brazing of zirconia. Acta Mater. 2000, 48, 4673–4676. [Google Scholar] [CrossRef]
  41. Zhu, J.; Kamiya, A.; Yamada, T.; Shi, W.; Naganuma, K.; Mukai, K. Surface tension, wettability and reactivity of molten titanium in Ti/yttria-stabilized zirconia system. Mater. Sci. Eng. A 2002, 327, 117–127. [Google Scholar] [CrossRef]
  42. Wang, N.; Wang, D.P.; Yang, Z.W.; Wang, Y. Interfacial microstructure and mechanical properties of zirconia ceramic and niobium joints vacuum brazed with two Ag-based active filler metals. Ceram. Int. 2016, 42, 12815–12824. [Google Scholar] [CrossRef]
  43. Zhang, Q.; Wang, J.; Zheng, K.; Lu, Y.; Yang, H. Microstructure evolution and mechanical properties of ZrO2/ZrO2 joints brazed with Ni–Ti filler metal. Mater. Res. Express 2019, 6, 126547. [Google Scholar] [CrossRef]
Figure 1. Microstructures of substrates and schematic diagram of brazing assembly. (a) Metallographic figure of α-Ti alloy, (b) XRD pattern of α-Ti alloy, (c) BSE image of ZrO2 ceramic and (d) brazing assembly.
Figure 1. Microstructures of substrates and schematic diagram of brazing assembly. (a) Metallographic figure of α-Ti alloy, (b) XRD pattern of α-Ti alloy, (c) BSE image of ZrO2 ceramic and (d) brazing assembly.
Metals 10 00526 g001
Figure 2. Typical microstructures and elemental distribution of titanium/Au/ZrO2 joint at 1150 °C for 10 min: (a) lower and (b) higher magnification of the typical interfacial microstructure, (c) line distributions and (dg) map distributions of Ti, Au, Zr and O.
Figure 2. Typical microstructures and elemental distribution of titanium/Au/ZrO2 joint at 1150 °C for 10 min: (a) lower and (b) higher magnification of the typical interfacial microstructure, (c) line distributions and (dg) map distributions of Ti, Au, Zr and O.
Metals 10 00526 g002
Figure 3. Magnification microstructure of the titanium/Au/ZrO2 joint: (a) the titanium/brazing seam interface; (b) the brazing seam/ZrO2 interface.
Figure 3. Magnification microstructure of the titanium/Au/ZrO2 joint: (a) the titanium/brazing seam interface; (b) the brazing seam/ZrO2 interface.
Metals 10 00526 g003
Figure 4. Microstructure of the titanium/Au/ZrO2 joint brazed at different brazing temperatures for 10 min: (a) 1110 °C, (b) 1130 °C, (c) 1150 °C, (d) 1170 °C and (e) 1190 °C.
Figure 4. Microstructure of the titanium/Au/ZrO2 joint brazed at different brazing temperatures for 10 min: (a) 1110 °C, (b) 1130 °C, (c) 1150 °C, (d) 1170 °C and (e) 1190 °C.
Metals 10 00526 g004
Figure 5. Microstructure of the TiO layer at different brazing temperature for 10 min: (a) 1110 °C, (b) 1130 °C, (c) 1150 °C, (d) 1170 °C and (e) 1190 °C.
Figure 5. Microstructure of the TiO layer at different brazing temperature for 10 min: (a) 1110 °C, (b) 1130 °C, (c) 1150 °C, (d) 1170 °C and (e) 1190 °C.
Metals 10 00526 g005
Figure 6. Microstructure of the titanium/Au/ZrO2 joint brazed at 1150 °C for different holding times: (a) 5 min, (b) 10 min, (c) 15 min, (d) 20 min and (e) 30 min.
Figure 6. Microstructure of the titanium/Au/ZrO2 joint brazed at 1150 °C for different holding times: (a) 5 min, (b) 10 min, (c) 15 min, (d) 20 min and (e) 30 min.
Metals 10 00526 g006
Figure 7. Effect of brazing parameters on shear strength of titanium/Au/ZrO2 joint. (a) Brazing temperature and (b) holding time.
Figure 7. Effect of brazing parameters on shear strength of titanium/Au/ZrO2 joint. (a) Brazing temperature and (b) holding time.
Metals 10 00526 g007
Figure 8. Fracture paths, high-magnification fractographs and XRD patterns of titanium/Au/ ZrO2 joints brazed at different parameters after the shear test. (a1–a3) 1110 °C for 10 min, (b1–b3) 1150 °C for 10 min, and (c1–c3) 1150 °C for 30 min.
Figure 8. Fracture paths, high-magnification fractographs and XRD patterns of titanium/Au/ ZrO2 joints brazed at different parameters after the shear test. (a1–a3) 1110 °C for 10 min, (b1–b3) 1150 °C for 10 min, and (c1–c3) 1150 °C for 30 min.
Metals 10 00526 g008
Table 1. Energy dispersive spectroscopy (EDS) results of the spots marked in Figure 3 (at.%).
Table 1. Energy dispersive spectroscopy (EDS) results of the spots marked in Figure 3 (at.%).
SpotTiAuZrOPossible Phases
A75.6722.480.021.83Ti3Au
B49.6047.670.062.67TiAu
C32.0561.520.036.40TiAu2
D18.4473.800.047.72TiAu4
E40.292.073.0254.62TiO

Share and Cite

MDPI and ACS Style

Lei, Y.; Bian, H.; Fu, W.; Song, X.; Feng, J.; Long, W.; Niu, H. Evaluation of Biomedical Ti/ZrO2 Joint Brazed with Pure Au Filler: Microstructure and Mechanical Properties. Metals 2020, 10, 526. https://doi.org/10.3390/met10040526

AMA Style

Lei Y, Bian H, Fu W, Song X, Feng J, Long W, Niu H. Evaluation of Biomedical Ti/ZrO2 Joint Brazed with Pure Au Filler: Microstructure and Mechanical Properties. Metals. 2020; 10(4):526. https://doi.org/10.3390/met10040526

Chicago/Turabian Style

Lei, Yuzhen, Hong Bian, Wei Fu, Xiaoguo Song, Jicai Feng, Weimin Long, and Hongwei Niu. 2020. "Evaluation of Biomedical Ti/ZrO2 Joint Brazed with Pure Au Filler: Microstructure and Mechanical Properties" Metals 10, no. 4: 526. https://doi.org/10.3390/met10040526

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop