Next Article in Journal
A Rational Design of a CoS2-CoSe2 Heterostructure for the Catalytic Conversion of Polysulfides in Lithium-Sulfur Batteries
Next Article in Special Issue
Study of the Relationship between Entropy and Hardness in Laser Cutting of Hardox Steel
Previous Article in Journal
Ballistic Impacts with Bullet Splash—Load History Estimation for .308 Bullets vs. Hard Steel Targets
Previous Article in Special Issue
Effect of Heat Input on Microstructure and Mechanical Properties of Deposited Metal of E120C-K4 High Strength Steel Flux-Cored Wire
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Advances in Processing of Titanium and Titanium Alloys through Metal Injection Molding for Biomedical Applications: 2013–2022

Department of Mechanical and Manufacturing Engineering, Faculty of Engineering and Built Environment, Universiti Kebangsaan Malaysia, Bangi 43600, Selangor, Malaysia
*
Author to whom correspondence should be addressed.
Materials 2023, 16(11), 3991; https://doi.org/10.3390/ma16113991
Submission received: 16 April 2023 / Revised: 14 May 2023 / Accepted: 23 May 2023 / Published: 26 May 2023
(This article belongs to the Special Issue Advances in Materials Processing (Second Volume))

Abstract

:
Metal injection molding (MIM) is one of the most widely used manufacturing processes worldwide as it is a cost-effective way of producing a variety of dental and orthopedic implants, surgical instruments, and other important biomedical products. Titanium (Ti) and Ti alloys are popular modern metallic materials that have revamped the biomedical sector as they have superior biocompatibility, excellent corrosion resistance, and high static and fatigue strength. This paper systematically reviews the MIM process parameters that extant studies have used to produce Ti and Ti alloy components between 2013 and 2022 for the medical industry. Moreover, the effect of sintering temperature on the mechanical properties of the MIM-processed sintered components has been reviewed and discussed. It is concluded that by appropriately selecting and implementing the processing parameters at different stages of the MIM process, defect-free Ti and Ti alloy-based biomedical components can be produced. Therefore, this present study could greatly benefit future studies that examine using MIM to develop products for biomedical applications.

Graphical Abstract

1. Introduction

The history and development of biomaterials are long and rich. Archaeologists have discovered that biomaterials have been used as dental implants since as early as 200 A.D. [1,2,3]. According to Vallet-Regí [4], after World War II, multiple materials were examined to create prostheses and implants for soldiers wounded in action. At that time, the term “biocompatibility” solely described how well an organism tolerated a material. However, in the 1960s, some of these implants led to complications. This created a need to identify implant materials that did not cause more problems than they effectively solved, as well as how to implant these new materials. In the late 1960s, surgical, medical, and dental laboratories began examining effective methods of producing replacement parts and implants for bodily injuries and published their findings in biomedical literature. Since then, the prevalence of such implants has increased globally.
Biomaterial implants are commonly used to replace and restore damaged or deteriorated organs or tissues in the human body [2]. This includes dental and orthopedic implants, ligaments, intraocular lenses, vascular grafts, artificial hearts, heart valves, biosensors, and cardiac pacemakers [5,6,7,8,9,10,11,12,13]. An implant should function flawlessly for a lifetime. As such, the development of materials that can withstand long-term implantation in the human body is a top priority, as, at present, commercially available biomaterials tend to break down after extended use due to poor biocompatibility, low fatigue strength, low corrosion and wear resistance, and a higher modulus than bone [14]. Biological biocompatibility, corrosion resistance, and mechanical biocompatibility determine the biocompatibility of orthopedic implant materials [7,15]. Biological biocompatibility examines interactions between an implant material and biological processes and includes its cancer-causing, mutational, genotoxic, or cytotoxic potential. Corrosion resistance in a biological environment is another important characteristic, as long-term orthopedic implants must be mechanically biocompatible and possess high wear resistance, high strength, and a low Young’s modulus [7]. The Young’s modulus is an important mechanical quality, as the effects of stress shielding often result in revision surgeries [15,16,17,18].
Biocompatible metals have so far been most widely used in biomedical applications. Titanium (Ti) and Ti alloys, stainless steel, and cobalt (Co) alloys are the broadly recognized biocompatible metals in the medical industry [19,20,21,22,23,24,25,26,27,28,29,30,31,32,33,34,35,36]. These materials are commonly used to replace and support fractured bone fragments as well as in dental implants, pacemaker casings, artificial heart valves, screws, plates, artificial joints, extrinsic fixators, spinal fixations, and stents [2,34,35,36].
Researchers currently prefer Ti and Ti alloys as the most advantageous biocompatible metals [19,20,21,22,23,24,25]. The utilization of Ti and Ti alloys as implant materials is deemed favorable due to their optimal properties [35]. As Ti and Ti alloys are chemically inert and possess strong fatigue resistance and a low Young’s modulus, they perform better than stainless steel and Co alloys in long-term implantation [21,37,38]. In addition, when compared to other biocompatible metals, Ti and Ti alloys exhibit significantly greater corrosion resistance [35,37]. Corrosion usually shortens the lifespan of the implants and necessitates additional surgery to replace the damaged ones.
Powder injection molding (PIM) is an economical method of mass-producing shaped components. Metal injection molding (MIM), a type of PIM, has made inroads into numerous industries as it is fast, cost effective, versatile, and potentially downsizable, as well as possessing outstanding design, dimensional accuracy, and mechanical properties that yield an exceptional surface finish and minimal by-products or waste [39,40,41,42,43,44,45,46,47,48]. At present, MIM is a popular method of manufacturing precise net-shaped parts for orthopedic and dental implants, surgical tools, and medical equipment [21,49,50,51,52,53,54,55]. Figure 1 depicts some commonly used MIM-fabricated medical tools and equipment. Aust et al. [56] used MIM to produce Ti-6Al-7Nb alloy bone screw implants to restore a fractured dens axis (Figure 2). Meanwhile, Barbosa et al. [57] and Barbosa [58] used two-component MIM to develop a Ti spinal implant (Figure 3). The study found that the highly porous section of the implant facilitated osseointegration via cell ingrowth and bodily fluid circulation in the interconnected pores, while the low-porosity section provided mechanical stability. Shu et al. [59] successfully fabricated MIM-based intravascular stents using 316L stainless steel powder in an effort to create stents with greater biocompatibility (Figure 4). The main steps of the production process included mixing, injection molding, debinding, and sintering (Figure 5). The binders were first mixed with the metal powder prior to undergoing the MIM process to produce a homogeneous feedstock, which was then converted into suitably sized pellets. During the MIM process, pressure and heat were used to mold the feedstock into the required shape. It was then demolded by removing the green part from the cavity of the mold. Solvent and thermal debinding techniques were used to remove the soluble and insoluble binders from the green part, respectively. Lastly, the brown debound part was sintered to produce the required mechanical and physical properties.
In the MIM process, it is highly important to get defect-free components with the desired mechanical properties. Choosing the appropriate process parameters at all stages of the MIM process is crucial to obtaining flawless components. Moreover, sintering is an indispensable step in the MIM process, and sintering parameters, especially sintering temperature, have a big impact on the mechanical properties of the components. The influence of sintering temperature on the microstructures and mechanical properties of different components is now being studied by MIM researchers from several domains [39,43,45,46].
In view of the above, as Ti and Ti alloys are superior to other existing biocompatible metals, this paper reviewed the MIM process parameters that studies published between 2013 and 2022 used to produce Ti and Ti alloy components for biomedical applications. In addition, this present study outlines the investigations of the MIM researchers over the past 10 years to attain desirable mechanical properties for Ti and Ti alloy-based biomedical components by using varying sintering temperatures.

2. Biomedical Applications of Titanium and Titanium Alloys

Ever since William Justin Kroll first discovered a method of extracting metallic Ti from its ore in the 1940s, Ti has increasingly been used in industrial and commercial applications [62,63]. Ti and Ti alloys have been extensively used in a variety of industrial fields, such as the aerospace, nuclear, marine, biomedical, automotive, and chemical industries, as they have high specific strength, low density, good biocompatibility, high fatigue strength, and good corrosion resistance [64,65,66,67,68,69,70,71,72,73,74,75]. Ti and Ti alloys provide good biocompatibility and superior corrosion performance as they naturally develop an uninterrupted, adhesive, and thin oxide film across the entire surface [76,77,78,79,80]. Figure 6 illustrates the two crystal orientations that can exist in Ti, an allotropic element. At temperatures below 882 °C, Ti remains in the alpha (α) phase with hexagonal, close-packed crystals. However, at temperatures above 882 °C, it transitions to the beta phase (β) and has body-centered cubic crystals [81,82,83]. Ti alloys with persistent α, β, and α + β phases can be produced at ambient temperature by selectively alloying them with other elements [81,84]. Aluminum (Al), carbon (C), oxygen (O), and nitrogen (N) stabilize Ti in the α phase, while manganese (Mn), chromium (Cr), iron (Fe), and vanadium (V) stabilize it in the β phase [82,85]. Hundreds of different types of Ti alloys have been developed worldwide, 20 to 30 of which, such as Ti-6Al-4V, Ti-2Al-2.5Zr, Ti-5Al-2.5Sn, Ti-Mo-Ni, Ti-32Mo, Ti-Pd, Ti-10-5-3, Ti-811, Ti-1023, Ti-1100, Ti-6242, IMI829, IMI834, BT9, and BT20, are very popular [86,87,88].
In the 1950s, Brånemark discovered that it was possible to permanently integrate Ti into bone. As such, they coined the term “osseointegration”, which refers to the secure fixation of Ti to bone tissues [89,90,91]. Osseointegration is commonly used to describe the biocompatible interactions that occur between a biomedical implant and the surrounding bone tissues that enable it to integrate with the bone tissues [92]. Several studies have examined the advantages of osseointegration in implants [93,94,95,96,97,98]. Sidambe [61] reported that Ti and Ti alloys have been extensively used as biomedical implants since the early 1970s. According to Manivasagam et al. [99], a combined total of GBP 2.2 million of Ti implants are transplanted into patients around the globe every year. Ti is commonly used as orthopedic and dental implants, artificial knee and hip joints, pacemakers, artificial hearts, cardiac valve prostheses, cornea backplates, bone plates, and screws for fracture fixation [61,99,100,101]. Stress shielding, which causes bone resorption, occurs when the Young’s moduli of the bone and the implant material are dissimilar [18,102,103,104]. The Young’s modulus of bone ranges between 10 and 30 GPa [18,105]. As seen in Table 1, the Young’s moduli of Ti alloys are more analogous to that of bone than other metal implant materials. Ti with a porosity of around 60% to 70% has outstanding Young’s modulus (Figure 7) and is, therefore, a promising material for bone implants [83,106]. However, the strength and porosity of a material must be balanced, as they both affect the efficacy of an implant [83,107,108].
The Ti-6Al-4V alloy, which was developed in the United States, was used for the first time in the medical industry in 1954 [2]. As Ti-6Al-4V possesses superior mechanical properties, is biocompatible, and does not interfere with innovative imaging methods such as computer tomography and magnetic resonance imaging, it remained the primary Ti alloy used in orthopedics and dentistry for a very long time [115,120,121,122,123,124,125,126,127,128,129,130]. However, multiple studies have raised the issue of the toxicity of the Al and V found in the Ti-6Al-4V alloy. Bodily exposure to Al and V has been found to cause cytotoxicity, allergic reactions, and neurological disorders [7,131,132,133,134,135]. Furthermore, although the Young’s modulus of Ti-6Al-4V (110 to 114 GPa) is lower than that of other biocompatible metals, such as stainless steel and Co alloys (Table 1), it is significantly higher than the stiffness of bone (10 to 30 GPa). Therefore, the mechanical loads of day-to-day tasks could cause stress shielding and lead to peri-prosthetic fractures, implant loosening, and bone loss [7,136,137,138,139]. As such, multiple studies have examined β-type Ti alloys to overcome the limitations of Ti-6Al-4V and other α + β-type Ti alloys [83,115,135]. According to Niinomi [140], β-type Ti alloys have low stiffness, which aids bone repair and remodeling. As β-type Ti alloys contain β-stabilizing elements, such as Nb, zirconium (Zr), molybdenum (Mo), and tantalum (Ta), their Young’s moduli are lower and closer to those of human bone, which could prevent implant loosening and bone resorption from stress-shielding. Furthermore, as they develop additional stable oxide layers, they have good biocompatibility, exceptional corrosion resistance, and are non-toxic [135,141]. As seen in Table 1, many types of β-type Ti alloys have been developed to achieve desirable Young’s modulus, tensile strength, yield strength, and elongation for biomedical applications. In this context, it is worthwhile to mention that the phase transformation process and mechanical properties of β-Ti alloys are significantly impacted by the injection molding, debinding, and sintering parameters during the MIM process.

3. MIM Process for Titanium and Titanium Alloys

3.1. Metallic Powder Selection

The attributes of a metallic powder, especially particle size and shape, substantively affect the quality of the feedstock as well as important stages of the MIM process [142,143,144,145,146,147,148,149]. According to multiple studies, metallic powder particles should possess the following characteristics to successfully complete the MIM process [39,150,151]:
  • Spherical-shaped powder particles are preferred, as irregular-shaped particles cause imbalanced powder loading and more shrinkage;
  • A higher packing density is required to maximally load the powder;
  • Adequate interparticle friction is required to retain the shape throughout the debinding stage. Distortions are more likely to occur when larger powder particles are used, as the intraparticle contact per unit volume decreases;
  • The powder particles must be non-agglomerated to produce defect-free sintered components;
  • The particles should not react with the multicomponent binder system;
  • The powder particles should be void-free to ensure that the sintered parts have excellent density.
As powder particle size significantly affects the MIM process, the right powder size must be chosen to produce flaw-free components. Most studies use particles that are less than 22 µm in size to improve densification, which affects the mechanical and corrosion properties of the product [2,151]. According to Piotter et al. [152], fine powder particles are preferable for metal materials; however, commercial-grade metal powders are larger than 5 µm on average. Table 2 lists the powder particle sizes of Ti and Ti alloys that MIM studies published between 2013 and 2022 used in biomedical applications.
As seen in Table 2, Ti and Ti-6Al-4V were most commonly used in MIM studies. According to Dehghan-Manshadi [179], commercially pure Ti is more commonly used in the MIM process as it is more tolerant of O, which may reach 0.4% in grade 4, than Ti-6Al-4V, which can only tolerate 0.2% O. The study found that MIM-processed, commercially pure Ti produces components with mechanical properties and chemical compositions that satisfy ASTM standards. Meanwhile, other studies have demonstrated that MIM-processed Ti-6Al-4V components have mechanical properties that are comparable to ASTM standards, especially when combined with small amounts of other elements, such as boron (B), C, gadolinium (Gd), and titanium carbide (TiC) [171,173,179,180,181,182,183,184,185,186]. As seen in Table 2, more studies have investigated the use of micro-sized powder particles in MIM than nanosized powder particles. This is because fine powder particles have a host of challenges, such as difficulty increasing the viscosity of the feedstock, difficulty attaining higher packing densities due to agglomeration, and more time needed to produce a homogeneous feedstock [39,187,188]. Furthermore, as seen in Table 2, only a few studies have examined the use of powder particles that are less than 22 µm in size. Larger particles decrease the debinding strength and increase the likelihood of distortion. Moreover, unlike smaller particles, larger particles have less intrinsic strength when they are packed together [151].
Powder particles are either spherical or irregular in shape, both of which possess unique properties. Spherical Ti powder particles that are less than 30 µm in size and produced using plasma atomization (from wire), gas atomization (from liquid), and plasma spheroidization (from non-spherical powder) are excellent for MIM as they possess excellent flow properties and shrink homogeneously throughout sintering [179]. Fine spherical Ti powder is expensive and has a low O tolerance, while non-spherical Ti powder costs significantly less and has a high O tolerance. As such, multiple studies have examined developing MIM procedures that can be used for hydride-dehydride (HDH) Ti powders [179,189,190,191,192,193]. Another strategy to lower the cost of MIM-fabricated components and increase the suitability of HDH powders for MIM is to combine HDH with spherical powders [193]. Figure 8 depicts the scanning electron microscopy (SEM) images of gas atomized, plasma atomized, and HDH Ti powders for MIM. As seen, the gas-atomized and advanced plasma-atomized Ti powders were spherical, while the HDH Ti powders were irregular and, therefore, a promising cost-effective alternative for MIM [68].

3.2. Binder Selection

A binder system is used in the MIM process to transport metal. The choice of binder is crucial, as it affects the quality of the finished product. The purpose of a binder system is to ensure that the shape of an MIM part remains unchanged until the sintering process begins as well as to help the formation during MIM. The choice of binder affects multiple variables and processes, such as packing density, powder-binder mix, feedstock flowability, injection molding, binder extraction, dimensional accuracy, and defect generation. Binders for the MIM process should possess the following characteristics [39,173,194]:
  • Outstanding adhesion to Ti powder;
  • Low likelihood of powder-binder separation;
  • Superior wetting capabilities and flow properties;
  • A viscosity that only varies slightly during MIM;
  • Produce very little residual C post-burnout;
  • Does not cause chemical reactions when interacting with Ti powder particles;
  • Toxicity-free.
The filler phase enables the feedstock to flow easily, while the backbone polymer serves as a second component that helps retain the shape of the sample. Additives, such as lubricants, surfactants, and dispersants, are commonly used as a final binder to improve the powder-binder interface. Table 3 lists the binders that MIM studies published between 2013 and 2022 used.
As seen in Table 3, most of these studies examined the use of paraffin wax for feedstock flowability, polyethylene as the backbone polymer for Ti and Ti alloys in biomedical applications, and stearic acid as a surfactant as it promotes effective powder-binder adhesion. More recent MIM studies, however, examined the use of palm stearin as a binder. Iriany et al. [196] were the first to examine the use of a palm stearin binder to circumvent debinding, which is the most time-consuming stage of the MIM process. The ability of this binder to serve as both a lubricant and a surfactant is a very important property [176,197]. Indonesia, Malaysia, Nigeria, and Colombia, as well as other African, Latin American, and Asian nations, are directly involved in palm oil production. This has made palm stearin more accessible to end users as a cost-effective commercial binder. Furthermore, as a palm stearin binder poses few environmental hazards, it is highly recommended for use in MIM [198].

3.3. Preparation of Feedstock

Ti and Ti alloys are typically mixed with binders to produce feedstocks. Powder loading and binder type significantly affect the characteristics of the produced feedstock. Optimal powder loading, based on critical powder loading, is recommended when using MIM to fabricate Ti and Ti alloy-based biomedical components. The powder particles must be in void-free contact with one another for critical powder loading [199,200]. Unlike critical powder loading, optimal powder loading uses fewer powder particles, significantly decreases feedstock viscosity, greatly simplifies the MIM process, and produces products with excellent mechanical properties and only minor distortions [201]. Higher powder loadings complicate the mixing process and increase feedstock viscosity [202]. On the other hand, higher binder content and insufficient powder loading in the feedstock cause higher shrinkage in sintered components. Lower powder content significantly affects the density of sintered components [203].
Powder particle size significantly affects critical powder loading during MIM. Smaller powder particles have a larger surface area, which lowers critical powder loading [204]. As such, many studies have examined multiple methods of determining the critical powder loading for MIM. Aggarwal et al. [205] used a torque rheometer to calculate the critical powder loading based on the rheological behavior of the feedstock. The powder loading range at which dramatic transformations in rheological parameters, such as viscosity, power law exponent, and flow activation energy, occurred was designated as the critical solid loading. Kong et al. [206] used a twin-screw mixer to gradually increase the volume of powder in the binder and a different strategy to determine the critical powder loading. The mixing torque was found to increase drastically when the critical powder loading was at a specific level. Mutsuddy and Ford [207], as well as Reddy et al. [208], gradually added an oil binder to a powder to obtain a critical powder volume concentration. The incremental addition of oil caused the torque to initially increase sharply before gradually decreasing until it reached a stable state. The critical powder loading was determined to be when the torque increased to its highest. Optimal powder loading is critical to producing defect-free MIM-processed Ti and Ti alloy components. This can be achieved by using powder loadings that are 2 to 5 vol.% less than the critical powder loading [199].
As seen in Table 4, most MIM studies used powder loadings of 50 and 67 vol.% for Ti and Ti alloys, while others used powder loadings of 70 vol.% and higher. De Freitas Daudt et al. [73] produced extremely porous MIM-processed Ti and reported that it was difficult to obtain open surface porosity and to retain the shape during debinding and sintering. The study examined the use of three different powder loadings: 72, 75, and 80 vol.%, and found that MIM components with 72 vol.% powder loading entirely collapsed (Figure 9b), 75 vol.% yielded less shape deformation (Figure 9c), and 80 vol.% powder loading yielded the best shape retention post-thermal debinding and sintering (Figure 9d).

3.4. Rheological Properties

The rheological characteristics of the feedstock, which closely correlate with viscosity, shear rate, and shear stress, can be analyzed to optimize the injection molding process. A capillary rheometer is typically used to measure rheological properties. At a specific molding temperature, a shear rate of 102 to 105 s−1 and a viscosity of less than 1000 Pa·s are best for Ti- and Ti alloy-based feedstocks to flow into the mold cavity [19,176,199]. The different rheological behaviors that a fluid exhibits when flowing are a critical factor. When shear thickening or dilatant flow occurs, the viscosity increases as the shear rate increases, which increases the likelihood of powder-binder segregation during MIM [200]. Conversely, when shear thinning or a pseudo-plastic flow occurs, the viscosity decreases as the shear rate increases, which is desirable for MIM as it helps fill the mold cavity effectively and faultlessly as well as maintain the shape of MIM parts [68,151,175,199]. Hayat et al. [157] produced MIM-processed Ti components using pure Ti powder by ensuring that the formulated feedstocks exhibited pseudo-plastic behavior during rheological analysis (Figure 10). According to Thavanayagam et al. [147], as some feedstocks that exhibit pseudo-plastic behavior may also exhibit dilatant flow behavior when subjected to high shear rates, it is necessary to thoroughly evaluate the rheological properties across a range of shear rates.
Equation (1) depicts the interdependence of viscosity η and shear rate Υ in non-Newtonian fluids [210]:
η = K Υ n 1
where K is a constant and n is the flow behavior index. In general, n provides a clear understanding of the shear sensitivity and flow behavior of the powder-binder mixture. When n > 1, the feedstock exhibits dilatant behavior. Conversely, the feedstock exhibits pseudo-plastic behavior when 𝑛 < 1 [176,177,210]. Temperature-dependent viscosity is another significant factor affecting the rheological behavior of a powder-binder mixture. Equation (2) depicts the flow activation energy E , which correlates with the Arrhenius equation and, generally, depicts how temperature affects feedstock viscosity [211]:
η T = η o exp   E / R T
where η o , R , and T are the viscosities at the reference temperature, gas constant, and temperature, respectively. E can be obtained from the slope of the plot of l n ( η ) against 1 / T . The smaller the E , the less susceptible the viscosity was to temperature changes, which prevented unpredictable flows throughout the MIM [202,212].
Rheological measurement is a pivotal step in MIM. Multiple studies have analyzed the rheology of Ti and Ti alloy feedstocks, with most of them conducting MIM with feedstocks that exhibit pseudo-plastic behavior [19,154,159,176,177]. An optimal powder loading and a feedstock that exhibits pseudo-plastic behavior could produce crack-free components.

3.5. Metal Injection Molding Process

The second step of MIM is to mold the prepared Ti and Ti alloy feedstocks into the desired shape. Molding parameters significantly affect the characteristics of the injected pieces, with defects frequently occurring during MIM. Defects that emerge during the final phases of MIM typically cannot be eliminated. Short shots, flashing, jetting, powder-binder segregation, and cracks are the most common defects that occur during MIM. While some of these flaws are easily observable, others can only be found during debinding and sintering. However, proper injection molding parameters, such as injection pressure, holding pressure, melt and mold temperature, and injection duration, can be used to correct these defects.
In the field of injection molding, a short shot, a defect that occurs when the powder loading, mold temperature, and injection pressure are incorrect, is a potential hindrance. Moghadam et al. [68] observed short-shot defects in injection-molded Ti components (Figure 11). As seen in Figure 11a–e, although the green parts were manufactured using injection processes at temperatures of 150 to 165 °C, short shot defects were only prevented when a feedstock temperature and powder loading of 165 °C and 53 vol.%, respectively, were used. As such, the study examined the best method of regulating feedstock temperature and powder loading to produce flaw-free components. Flashing is another common defect that occurs during injection molding. Urtekin et al. [19] produced MIM-processed cortical-bone screws with Ti-6Al-4V powders and binders and reported that the green component maintained its geometry at a pressure of 120 MPa (Figure 12). As seen in Figure 12, challenges, including ejection from mold and flashing, occurred when the pressure exceeded 120 MPa. These issues were believed to arise due to high injection pressures. Therefore, flash-free components could be produced by using the correct injection pressure and properly clamping the mold during injection molding.
In summary, it is crucial to obtain defect-free green Ti and Ti alloy components during injection molding. As such, the injection molding parameters should be correctly adjusted to produce flawless parts for debinding and sintering.

3.6. Debinding Process

Solvent and thermal debinding are debinding methods that are frequently used in MIM studies. Solvent debinding, which involves immersing injection-molded components in a liquid solvent and conducting the procedure at a specific temperature, can be used to extract soluble binders from a multi-component binder system [39,73,170]. Diffusion is commonly used to eliminate the soluble binder. Yang et al. [213] schematically outline the various stages of the solvent extraction process (Figure 13 and Figure 14). As seen in Figure 13 and Figure 14, the diffusion of solvent molecules into the binder marks the beginning of the solvent extraction process, which causes a swelling gel to form. The soluble binder dissolves into the solution due to sufficiently potent interactions between the solvent and the soluble binder that counteract the intermolecular forces [214]. The pore size increases as the insoluble binder is kept in the contact area. This open-pore structure accelerates the thermal extraction process and facilitates the quick removal of the insoluble binder [215]. Table 5 lists the parameters that extant studies have used to solvent debind a variety of injection-molded Ti and Ti alloy components.
As seen in Table 5, solvent debinding was mostly conducted at temperatures between 40 and 60 °C, and heptane and hexane were the preferred choices of solvent. Temperature is crucial in solvent debinding, as higher temperatures have been found to improve soluble binder and solvent interactions by transforming the solubility and the binder diffusion coefficient [216]. Often, very little soluble binder remains in components after solvent debinding. During the thermal debinding process, this can be eliminated along with the insoluble binder. It is important to produce solvent-debound components that are crack-free. As seen in Table 5, Moghadam et al. [68] examined the use of three different solvent debinding temperatures of 45, 60, and 75 °C on injection-molded Ti components or green parts. Although a temperature of 75 °C removed the binder very quickly, the solvent debinding process was halted immediately after 60 min as cracks began to form (Figure 15). This is believed to have occurred due to the extremely rapid extraction rate and the potential softening of the low-density polyethylene backbone. Components that were solvent debound at 45 and 60 °C were defect-free. These results indicate the importance of debinding temperature in solvent debinding. Therefore, defect-free solvent debound components can be produced by using the correct solvent type, solvent debinding temperature, and time.
Thermal debinding is commonly used to eliminate insoluble binders from injection-molded components. In the early stages of thermal debinding, the binder degradation rate determines how quickly the binder flows from inside the component to its surface. As the polymer degradation process progresses, pores near the surface open and help dispel the burnout gases of the binder via interconnected pore channels [199,217]. Due to polymer losses during thermal debinding, the component typically weakens immediately. As such, thermally debound components must be handled with care. According to Hamidi et al. [2], thermally debound components experience thermal, gravitational, and residual stresses, which could cause defects or cracks to form as the polymer degrades. Furthermore, the subsequent sintering process will aggravate the microscopic defects that formed during thermal debinding. As such, the thermal debinding schedule must be designed properly to overcome these issues. Table 6 lists the parameters that extant studies have used to thermally debind Ti and Ti alloy components. As seen in Table 6, temperatures between 500 and 700 °C were the best for the given thermal extraction durations.

3.7. Sintering Process

Sintering, the last stage of the MIM process, transforms debound components into a robust mass and significantly affects the physical and mechanical properties of the produced component. The sintering temperatures are typically 70 to 90% of the melting temperature of the metal [2,218]. Vacuum, air, argon, hydrogen, and reduced nitrogen/hydrogen atmospheres are commonly used during sintering. The exterior surface of the powder particles undergoes remarkable chemical changes and atomic diffusion when the sintering temperature increases from one-half to two-thirds of the melting point of the metal. Thermolysis occurs as the temperature increases, which burns out the remaining binders [2,39]. Sintering uses several mechanisms, such as lattice diffusion, surface diffusion, grain boundary diffusion, plastic flow, evaporation-condensation, and viscous flow, for mass transport [219]. These mechanisms encourage the growth of necks between particles, which improve the strength of the consolidated powders [39,150]. Table 7 lists the parameters that extant studies published between 2013 and 2022 have used to sinter Ti and Ti alloy components. As seen in Table 7, a temperature range of 1100 to 1500 °C, a heating rate of 3 to 10 °C/min, a duration of 1 to 8 h, and a vacuum or argon atmosphere were preferred.
It is important to mention that Ti and Ti alloys have a profound affinity for impurities, notably O and C [193]. O reduces the stress corrosion resistance, fatigue strength, and tensile ductility of the MIM-processed sintered Ti and Ti alloy components [179]. According to Luo et al. [220], the ultimate tensile strength and yield strength of the unalloyed Ti increased, but tensile ductility decreased as the O content increased (Figure 16). The initial powder, binder system, and sintering environment are the primary contributors to O contamination in the MIM process. Usually, with careful selection of powder and binder as well as precise control over the injection, debinding, and sintering processes, O content can be kept lower than 0.3 wt.% for commercially pure Ti components [179]. On the other hand, in order to prevent the development of TiC in structure, which could reduce corrosion resistance, weaken the elongation and fatigue properties, and increase the Young’s modulus of sintered parts, the C content must be kept to less than 0.08 wt.% [221].
During the MIM process, it is essential that sintered components exhibit the desired mechanical properties. Mechanical properties are significantly influenced by sintering parameters, particularly sintering temperature. Table 8 summarizes the mechanical properties of the Ti and Ti alloy components that sintered at different temperatures between 2013 and 2022.
Based on Table 8, a sintering temperature range of 1100 to 1500 °C was employed by the researchers to obtain a Young’s modulus of 7.80 to 103 GPa and a tensile strength of 395 to 1154 MPa. Dehghan-Manshadi et al. [156] examined the optimal MIM parameters for porous Ti scaffolds. Figure 17 depicts the SEM micrographs of components sintered at 1250 °C. However, the sintered components contained a few big and irregular-shaped pores that were 150 to 200 µm in size (Figure 17a) as well as microsized pores (Figure 17b) due to the elimination of the space holder and binder as well as the sintering of the Ti particles. The optimal pore size for body fluid transportation and mineralized bone growth ranges between 100 and 300 µm [222,223]. The study found that the rough interior walls of the big pores were better suited for new bone tissue ingrowth. An EDS of the sintered components revealed that the microstructure did not contain chlorine or potassium (Figure 17c); therefore, all the space holders had been completely removed during water immersion. Furthermore, components that had been sintered at 1250 °C had a Young’s modulus of 7.80 GPa. The Young’s modulus increased to 22.0 GPa at 1300 °C when the porosity decreased. Bootchai et al. [160] found that MIM-processed Ti components sintered at 1150 °C had a Young’s modulus and tensile strength of 99 GPa and 542 MPa, respectively. A close tensile strength of 617 MPa was obtained by Hayat et al. [157], who sintered Ti components at 1300 °C.
Dehghan-Manshadi et al. [46] identified the best processing parameters to fabricate MIM-processed HDH Ti with minor deformations. Spherically dispersed and small-sized pores were observed in the sintered sample (Figure 18a). A closer examination revealed that an almost non-porous, thin surface layer had formed (Figure 18b) in contrast to the central section of the component (Figure 18c). As seen in Figure 18d, the number of pores significantly decreased and the average pore size increased as the sintering temperature increased. Furthermore, changes in pore size and distribution due to increasing the sintering temperature may directly affect the mechanical properties of sintered components. Components sintered at 1250 °C had a tensile strength and elongation of 395 MPa and 12.5%, respectively.
As a prospective low-cost material, Santos et al. [165] fabricated Ti-Mn alloy components through MIM that had been sintered at 1100 °C for biomedical applications. Figure 19 depicts the optical micrographs of Ti-Mn alloys containing 8 to 17 mass% of Mn. The average grain diameter, which was 69.8 ± 6.1 μm, did not significantly differ between the alloys. As seen in Figure 19, each alloy contained large, interconnected pores and small, closed pores, with most of the pores (ellipses) located at the grain boundaries. An evaluation of the mechanical properties of the examined alloys indicated that the Ti-9Mn alloy had the best combination of tensile strength (1046 MPa) and elongation (4.7%). Figure 20 depicts the SEM fractographs of the tensile-tested components. As seen in Figure 20a–c, the Ti-9Mn and Ti-12Mn alloys had larger regions of concentrated cleavage-like structures and dimples than the Ti-8Mn alloy. Figure 20e,f depict the unique massive trans-granular cleavage-like structures found in the Ti-15Mn and Ti-17Mn alloys. The study also found that the Ti-8Mn to Ti-17Mn alloys had higher Young’s moduli. The Young’s moduli of the Ti-8Mn, Ti-9Mn, Ti-12Mn, Ti-13Mn, Ti-15Mn, and Ti-17Mn alloys were 87, 89, 96, 99, 98, and 103 GPa, respectively.
The mechanical properties of Ti-Nb alloys make them viable bioactive implants. Zhao et al. [167] fabricated MIM-processed Ti-16Nb alloys and observed microstructural changes in components sintered at four different temperatures: 900, 1100, 1300, and 1500 °C (Figure 21). As seen in Figure 21, the component sintered at 900 °C contained Ti and Nb particles that matched their primeval particle morphologies. Although Ti particle diffusion had begun, boundaries were clearly visible between the Ti and Nb particles. It was difficult to observe the Ti particle boundaries in the Ti-16Nb alloy sintered at 1100 °C as it mostly contained irregular-shaped residual pores. Unlike Ti-16Nb alloys sintered at 900 and 1100 °C, the removal of observable Nb particles and the existence of thin needle-shaped precipitates were the most clearly observable microstructural characteristics in Ti-16Nb alloys sintered at 1300 and 1500 °C. The porosity was also found to decrease from roughly 25 to 6% as the sintering temperature increased. The ultimate tensile strength and the Young’s modulus of the Ti-16Nb alloy sintered at 1500 °C were 667 MPa and 80 GPa, respectively. Bidaux et al. [169] sintered Ti-17Nb alloys at 1400 °C that had a Young’s modulus of 76 GPa. In comparison to Zhao et al. [167] and Bidaux et al. [169], a higher Young’s modulus of 100 GPa was obtained by Yılmaz et al. [168], who carried out the sintering of MIM-processed Ti-Nb alloys at 1500 °C.
Nor et al. [174] fabricated Ti-6Al-4V components on the grounds of MIM. The optimum sintering temperature was determined to be 1200 °C. Ti-6Al-4V alloys sintered at 1200 °C had tensile strengths of 934.33 MPa. Holm et al. [171] fabricated Ti-6Al-4V components with Gd powder and sintered them at 1350 °C. Adding Gd was found to increase the porosity of the Ti-6Al-4V alloys from 3.6 to 5%. However, the addition of 1 wt.% of Gd decreased tensile strength from 824 to 749 MPa. Therefore, Ti-6Al-4V + Gd components had higher porosity but lower tensile strength than Ti-6Al-4V components. Arifin et al. [176] examined the viability of fabricating PIM-processed Ti-6Al-4V and hydroxyapatite (HA) composites sintered at 1100, 1200, and 1300 °C. Figure 22 depicts the components sintered at different temperatures. As seen in Figure 22, residual HA particles were detected between the gaps of the Ti particles, and the Ti particles were covered in partially decomposed HA. Some HA groups were observed on the surface at 1100 °C; however, they were not detected when the sintering temperature increased significantly. The presence of Ti atoms accelerates the formation of titanium oxide (TiO2) and HA dehydroxylation [224,225]. Inter-diffusion occurred at the Ti-HA interface due to the migration of HA towards Ti bulk and Ti atoms to O and HA atoms. As O is an interstitial atom, it diffuses into the Ti lattice until saturation. The Ti was then oxidized, which slowed O diffusion. Composites sintered at 1300 °C had a Young’s modulus of 44.26 GPa, which is very close to that of human bone. Ramli et al. [177] reported that Ti-6Al-4V and bioactive wollastonite (WA) were excellent for implantation applications. The PIM-processed Ti6Al4V/WA composites sintered at 1100, 1200, and 1300 °C had Young’s moduli of 18.10, 15.62, and 14.57 GPa, respectively.
Kafkas and Ebel [113] fabricated Ti-24Nb-4Zr-8Sn components based on MIM. Components sintered at 1400 °C yielded the highest tensile strength (656 MPa) and Young’s modulus (54 GPa). Increasing the sintering temperature from 1400 to 1500 °C did not significantly affect strength. On the other hand, Suwanpreecha et al. [55] fabricated novel Ti-27.5Nb-8.5Ta-3.5Mo-2.5Zr-5Sn components sintered at 1000 to 1400 °C. Sintering at 1000 °C did not adequately sinter the components, while sintering at 1100 °C yielded the best tensile strength (1154 MPa) and the highest elastic modulus (98 GPa).

4. Concluding Remarks and Future Directions

MIM is an exemplary manufacturing approach for fabricating small and intricate components for the biomedical industry at a low cost. This present study reviewed the MIM parameters that extant studies have used to fabricate Ti and Ti alloy biomedical products over the past 10 years. Ti and Ti alloys outperform other biocompatible metals, such as stainless steel and Co alloys, in long-term implantation due to their low Young’s modulus, strong fatigue resistance, and chemical inertness. The drawbacks of the Ti-6Al-4V and other α + β-type Ti alloys have been mostly resolved by β-type Ti alloys. Due to containing β-stabilizing elements such as Zr, Nb, Mo, and Ta, the Young’s moduli of β-type Ti alloys are closer to those of human bone, which could impede the loosening of implants and the resorption of bone from stress-shielding. Moreover, Ti with a porosity between 60 and 70% is regarded as a viable material for bone implants due to its remarkable Young’s modulus. In this review paper, we acknowledge the extensive usage of Ti and Ti alloys as biomedical implants.
A crucial step in the MIM process is choosing the appropriate powder particle size. Although most studies used microsized instead of nanosized powder particles for Ti and Ti alloys, further research using fine particles could be useful to comprehend the impact of such particles on different stages of the MIM process. When preparing feedstock, the majority of researchers chose to employ high powder loading, but this often makes the process of mixing powder and binder challenging and is responsible for the incomplete filling of the mold cavity during the injection molding process. Therefore, it is recommended to use optimal powder loading. The MIM process is dramatically streamlined by optimal powder loading since it requires fewer powder particles compared to critical powder loading, substantially reduces the viscosity of feedstock, and yields products with superior mechanical properties and minimal deformations. Cracks, short shots, jetting, and flashing are common defects for injection-molded components. Some of these imperfections can be seen instantly, but others are only revealed during the debinding and sintering processes. Defects that arise during the injection process can be avoided by selecting the injection molding parameters appropriately. The debinding process is known to be a sensitive stage, as parts remain extremely fragile during this period. It is preferable not to employ an excessively high solvent debinding temperature during the solvent debinding process to achieve crack-free components. This review paper demonstrated the competence of MIM researchers to produce sintered Ti and Ti alloy biomedical components. Basically, the sintering temperature has an immense effect on the mechanical properties of the sintered components. Finally, a doorway into the future of biomedical applications may be opened by the production of multi-functional biomedical products using Ti and Ti alloy materials along with other metal- or ceramic-based materials through the two-component metal injection molding (2C-MIM) process.

Author Contributions

Conceptualization and methodology, A.B., N.M., A.B.S., N.H.J. and F.M.F.; writing—original draft preparation, A.B.; writing—review and editing, A.B., A.B.S., N.H.J. and F.M.F.; supervision, N.M. and A.B.S.; funding acquisition, N.M. and A.B.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Universiti Kebangsaan Malaysia grant number DIP-2022-015.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors would like to thank Universiti Kebangsaan Malaysia for the financial support under the grant number DIP-2022-015.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Todros, S.; Todesco, M.; Bangno, A. Biomaterials and their biomedical applications: From replacement to regeneration. Processes 2021, 9, 1949. [Google Scholar] [CrossRef]
  2. Hamidi, M.F.F.A.; Harun, W.S.W.; Samykano, M.; Ghani, S.A.C.; Ghazalli, Z.; Ahmad, F.; Sulong, A.B. A review of biocompatible metal injection moulding process parameters for biomedical applications. Mater. Sci. Eng. C 2017, 78, 1263–1276. [Google Scholar] [CrossRef] [PubMed]
  3. Malekani, J.; Schmutz, B.; Gu, Y.; Schuetz, M.; Yarlagadda, P. Biomaterials in orthopedic bone plates: A review. In Proceedings of the Annual International Conference on Materials Science, Metal and Manufacturing, Singapore, 12–13 December 2011. [Google Scholar]
  4. Vallet-Regí, M. Evolution of biomaterials. Front. Mater. 2022, 9, 864016. [Google Scholar] [CrossRef]
  5. Istrate, B.; Munteanu, C.; Antoniac, I.-V.; Lupescu, S.-C. Current research studies Mg-Ca-Zn biodegradable alloys used as orthopedic implant—Review. Crystals 2022, 12, 1468. [Google Scholar] [CrossRef]
  6. Schmidt, C.E.; Baier, J.M. Acellular vascular tissues: Natural biomaterials for tissue repair and tissue engineering. Biomaterials 2000, 21, 2215–2231. [Google Scholar] [CrossRef] [PubMed]
  7. Biesiekierski, A.; Wang, J.; Gepreel, M.A.-H.; Wen, C. A new look at biomedical Ti-based shape memory alloys. Acta Biomater. 2012, 8, 1661–1669. [Google Scholar] [CrossRef] [PubMed]
  8. Ratner, B.D.; Hoffman, A.S.; Schoen, F.J.; Lemons, J.E. Biomaterials Science: An Introduction to Materials in Medicine; Elsevier: Amsterdam, The Netherlands, 2004. [Google Scholar]
  9. Seemann, A.; Akbaba, S.; Buchholz, J.; Türkkan, S.; Tezcaner, A.; Woche, S.K.; Guggenberger, G.; Kirschning, A.; Dräger, G. RGD-modified titanium as an improved osteoinductive biomaterial for use in dental and orthopedic implants. Bioconjugate Chem. 2022, 33, 294–300. [Google Scholar] [CrossRef]
  10. Roman, A.-M.; Voiculescu, I.; Cimpoesu, R.; Istrate, B.; Chelariu, R.; Cimpoesu, N.; Zegan, G.; Panaghie, C.; Lohan, N.M.; Axinte, M.; et al. Microstructure, shape memory effect, chemical composition and corrosion resistance performance of biodegradable FeMnSi-Al alloy. Crystals 2023, 13, 109. [Google Scholar] [CrossRef]
  11. Lin, R.; Wang, Z.; Li, Z.; Gu, L. A two-phase and long-lasting multi-antibacterial coating enables titanium biomaterials to prevent implants-related infections. Mater. Today Bio 2022, 15, 100330. [Google Scholar] [CrossRef]
  12. Soykan, C. A theoretical approach to the structural, elastic and electronic properties of Ti8-xV4-yMox+y+zAl4-z lightweight shape memory alloys for biomaterial implant applications. Phys. B Condens. Matter 2020, 598, 412416. [Google Scholar] [CrossRef]
  13. Trincă, L.C.; Mareci, D.; Solcan, C.; Fântânariu, M.; Burtan, L.; Hriţcu, L.; Chiruţă, C.; Fernández-Mérida, L.; Rodríguez-Raposo, R.; Santana, J.J.; et al. New Ti-6Al-2Nb-2Ta-1Mo alloy as implant biomaterial: In vitro corrosion and in vivo osseointegration evaluations. Mater. Chem. Phys. 2020, 240, 122229. [Google Scholar] [CrossRef]
  14. Gepreel, M.A.-H.; Niinomi, M. Biocompatibility of Ti-alloys for long-term implantation. J. Mech. Behav. Biomed. Mater. 2013, 20, 407–415. [Google Scholar] [CrossRef] [PubMed]
  15. Hussein, A.H.; Gepreel, M.A.-H.; Gouda, M.K.; Hefnawy, A.M.; Kandil, S.H. Biocompatibility of new Ti-Nb-Ta base alloys. Mater. Sci. Eng. C 2016, 61, 574–578. [Google Scholar] [CrossRef] [PubMed]
  16. Moyer, J.A.; Metz, C.M.; Callaghan, J.J.; Hennessy, D.W.; Liu, S.S. Durability of second-generation extensively porous-coated stems in patients age 50 and younger. Clin. Orthop. Relat. Res. 2010, 468, 448–453. [Google Scholar] [CrossRef] [PubMed]
  17. Lenthe, G.H.V.; de Waal Malefijt, M.C.; Huiskes, R. Stress shielding after total knee replacement may cause bone resorption in the distal femur. J. Bone Jt. Surg. 1997, 79, 117–122. [Google Scholar] [CrossRef]
  18. Arifin, A.; Sulong, A.B.; Muhamad, N.; Syarif, J.; Ramli, M.I. Material processing of hydroxyapatite and titanium alloy (HA/Ti) composite as implant materials using powder metallurgy: A review. Mater. Des. 2014, 55, 165–175. [Google Scholar] [CrossRef]
  19. Urtekin, L.; Bozkurt, F.; Çanli, M. Analysis of biomedical titanium implant green parts by X-Ray tomography. J. Mater. Res. Technol. 2021, 14, 277–286. [Google Scholar] [CrossRef]
  20. Shanmuganantha, L.; Khan, M.U.A.; Sulong, A.B.; Ramli, M.I.; Baharudin, A.; Ariffin, H.M.; Razak, S.I.A.; Ng, M.H. Characterization of titanium ceramic composite for bone implants applications. Ceram. Int. 2022, 48, 22808–22819. [Google Scholar] [CrossRef]
  21. Rodriguez-Contreras, A.; Punset, M.; Calero, J.A.; Gil, F.J.; Ruperez, E.; Manero, J.M. Powder metallurgy with space holder for porous titanium implants: A review. J. Mater. Sci. Technol. 2021, 76, 129–149. [Google Scholar] [CrossRef]
  22. Milone, D.; Fiorillo, L.; Alberti, F.; Cervino, G.; Filardi, V.; Pistone, A.; Cicciù, M.; Risitano, G. Stress distribution and failure analysis comparison between zirconia and titanium dental implants. Procedia Struct. Integr. 2022, 41, 680–691. [Google Scholar] [CrossRef]
  23. Stricker, A.; Bergfeldt, T.; Fretwurst, T.; Addison, O.; Schmelzeisen, R.; Rothweiler, R.; Nelson, K.; Gross, C. Impurities in commercial titanium dental implants—A mass and optical emission spectrometry elemental analysis. Dent. Mater. 2022, 38, 1395–1403. [Google Scholar] [CrossRef]
  24. Mühl, A.; Szabó, P.; Krafcsik, O.; Aigner, Z.; Kopniczky, J.; Nagy, Á.; Marada, G.; Turzó, K. Comparison of surface aspects of turned and anodized titanium dental implant, or abutment material for an optimal soft tissue integration. Heliyon 2022, 8, e10263. [Google Scholar] [CrossRef]
  25. Chen, M.; Sun, Y.; Hou, Y.; Luo, Z.; Li, M.; Wei, Y.; Chen, M.; Tan, L.; Cai, K.; Hu, Y. Constructions of ROS-responsive titanium-hydroxyapatite implant for mesenchymal stem cell recruitment in peri-implant space and bone formation in osteoporosis microenvironment. Bioact. Mater. 2022, 18, 56–71. [Google Scholar] [CrossRef] [PubMed]
  26. Kheder, W.; Kawas, S.A.; Khalaf, K.; Samsudin, A.R. Impact of tribocorrosion and titanium particles release on dental implant complications—A narrative review. Jpn. Dent. Sci. Rev. 2021, 57, 182–189. [Google Scholar] [CrossRef] [PubMed]
  27. Alizade, C.; Jafarov, A.; Berchenko, G.; Bicer, O.S.; Alizada, F. Investigation of the process intergrowth of bone tissue into the hole in titanium implants (Experimental research). Injury 2022, 53, 2741–2748. [Google Scholar] [CrossRef] [PubMed]
  28. Liu, X.; Li, S.; Meng, Y.; Fan, Y.; Liu, J.; Shi, C.; Ren, F.; Wu, L.; Wang, J.; Sun, H. Osteoclast differentiation and formation induced by titanium implantation through complement C3a. Mater. Sci. Eng. C 2021, 122, 111932. [Google Scholar] [CrossRef]
  29. Zhang, S.; Yu, Y.; Wang, H.; Ren, L.; Yang, K. Study on mechanical behavior of Cu-bearing antibacterial titanium alloy implant. J. Mech. Behav. Biomed. Mater. 2022, 125, 104926. [Google Scholar] [CrossRef]
  30. Yan, B.; Tan, J.; Wang, D.; Qiu, J.; Liu, X. Surface alloyed Ti–Zr layer constructed on titanium by Zr ion implantation for improving physicochemical and osteogenic properties. Prog. Nat. Sci. Mater. Int. 2020, 30, 635–641. [Google Scholar] [CrossRef]
  31. Dewidar, M. Influence of processing parameters and sintering atmosphere on the mechanical properties and microstructure of porous 316L stainless steel for possible hard-tissue applications. Int. J. Mech. Mechatron. Eng. 2012, 12, 10–24. [Google Scholar]
  32. Fonseca-García, A.; Pérez-Alvarez, J.; Barrera, C.C.; Medina, J.C.; Almaguer-Flores, A.; Sánchez, R.B.; Rodil, S.E. The effect of simulated inflammatory conditions on the surface properties of titanium and stainless steel and their importance as biomaterials. Mater. Sci. Eng. C 2016, 66, 119–129. [Google Scholar] [CrossRef]
  33. Ibrahim, M.Z.; Sarhan, A.A.D.; Kuo, T.Y.; Yusof, F.; Hamdi, M. Characterization and hardness enhancement of amorphous Fe-based metallic glass laser cladded on nickel-free stainless steel for biomedical implant application. Mater. Chem. Phys. 2019, 235, 121745. [Google Scholar] [CrossRef]
  34. Muley, S.V.; Vidvans, A.N.; Chaudhari, G.P.; Udainiya, S. An assessment of ultra fine grained 316L stainless steel for implant applications. Acta Biomater. 2016, 30, 408–419. [Google Scholar] [CrossRef] [PubMed]
  35. Jiménez-Marcos, C.; Baltatu, M.S.; Florido-Suárez, N.R.; Socorro-Perdomo, P.P.; Vizureanu, P.; Mirza-Rosca, J.C. Mechanical properties and corrosion resistance of two new titanium alloys for orthopaedics applications. Mater. Today Proc. 2023, 72, 544–549. [Google Scholar] [CrossRef]
  36. Abdulah, N.; Omar, M.A.; Jamaludin, S.B.; Zainon, N.M.; Roslani, N.; Meh, B.; Jalil, M.N.A.; Hijazi, M.B.M.; Omar, A.Z. Innovative metal injection molding (MIM) method for producing CoCrMo alloy metallic prosthesis for orthopedic applications. Adv. Mater. Res. 2014, 879, 102–106. [Google Scholar] [CrossRef]
  37. Yemisci, I.; Mutlu, O.; Gulsoy, N.; Kunal, K.; Atre, S.; Gulsoy, H.O. Experimentation and analysis of powder injection molded Ti10Nb10Zr alloy: A promising candidate for electrochemical and biomedical application. J. Mater. Res. Technol. 2019, 8, 5233–5245. [Google Scholar] [CrossRef]
  38. Raabe, D.; Sander, B.; Friák, M.; Ma, D.; Neugebauer, J. Theory-guided bottom-up design of β-titanium alloys as biomaterials based on first principles calculations: Theory and experiments. Acta Mater. 2007, 55, 4475–4487. [Google Scholar] [CrossRef]
  39. Basir, A.; Sulong, A.B.; Jamadon, N.H.; Muhamad, N.; Emeka, U.B. Process parameters used in macro/micro powder injection molding: An overview. Met. Mater. Int. 2021, 27, 2023–2045. [Google Scholar] [CrossRef]
  40. Ferreira, T.J.; -Fernandes, C.P.; Piedade, A.P.; Tondela, J.; Vieira, M.T. Mechanical behaviour of dental implants manufactured from metallic powders by µMIM. Cienc. Tecnol. Mater. 2014, 26, 89–95. [Google Scholar]
  41. Zhou, C.; Jiang, H.; Liu, C.; Yan, B.; Yan, P. Microstructure features of high performance soft magnetic alloy Fe-3 wt.% Si prepared by metal injection molding. Mater. Chem. Phys. 2021, 273, 125068. [Google Scholar] [CrossRef]
  42. Langlais, D.; Demers, V.; Brailovski, V. Rheology of dry powders and metal injection molding feedstocks formulated on their base. Powder Technol. 2022, 396, 13–26. [Google Scholar] [CrossRef]
  43. Subaşı, M. The investigation of sintering temperature and Ni interlayer effects on diffusion bonding in inserted metal injection molding. J. Manuf. Process. 2020, 58, 706–711. [Google Scholar] [CrossRef]
  44. Askari, A.; Momeni, V. Rheological investigation and injection optimization of Fe-2Ni-2Cu feedstock for metal injection molding process. Mater. Chem. Phys. 2021, 271, 124926. [Google Scholar] [CrossRef]
  45. Mukund, B.N.; Hausnerova, B. Variation in particle size fraction to optimize metal injection molding of water atomized 17–4PH stainless steel feedstocks. Powder Technol. 2020, 368, 130–136. [Google Scholar] [CrossRef]
  46. Dehghan-Manshadi, A.; StJohn, D.; Dargusch, M.; Chen, Y.; Sun, J.F.; Qian, M. Metal injection moulding of non-spherical titanium powders: Processing, microstructure and mechanical properties. J. Manuf. Process. 2018, 31, 416–423. [Google Scholar] [CrossRef]
  47. Thavanayagam, G.; Swan, J.E. Aqueous debinding of polyvinyl butyral based binder system for titanium metal injection moulding. Powder Technol. 2018, 326, 402–410. [Google Scholar] [CrossRef]
  48. Basir, A.; Sulong, A.B.; Jamadon, N.H.; Muhamad, N. Feedstock properties and debinding mechanism of yttria-stabilized zirconia/stainless steel 17-4PH micro-components fabricated via two-component micro-powder injection molding process. Ceram. Int. 2021, 47, 20476–20485. [Google Scholar] [CrossRef]
  49. Torres, Y.; Lascano, S.; Bris, J.; Pavón, J.; Rodriguez, J.A. Development of porous titanium for biomedical applications: A comparison between loose sintering and space-holder techniques. Mater. Sci. Eng. C 2014, 37, 148–155. [Google Scholar] [CrossRef]
  50. Gemelli, E.; de Jesus, J.; Camargo, N.H.A.; de Almeida Soares, G.D.; Henriques, V.A.R.; Nery, F. Microstructural study of a titanium-based biocomposite produced by the powder metallurgy process with TiH2 and nanometric β-TCP powders. Mater. Sci. Eng. C 2012, 32, 1011–1015. [Google Scholar] [CrossRef]
  51. Liu, Y.; Li, K.; Luo, T.; Song, M.; Wu, H.; Xiao, J.; Tan, Y.; Cheng, M.; Chen, B.; Niu, X.; et al. Powder metallurgical low-modulus Ti–Mg alloys for biomedical applications. Mater. Sci. Eng. C 2015, 56, 241–250. [Google Scholar] [CrossRef] [PubMed]
  52. Wermuth, D.P.; Paim, T.C.; Bertaco, I.; Zanatelli, C.; Naasani, L.I.S.; Slaviero, M.; Driemeier, D.; Tavares, A.C.; Martins, V.; Escobar, C.F.; et al. Mechanical properties, in vitro and in vivo biocompatibility analysis of pure iron porous implant produced by metal injection molding: A new eco-friendly feedstock from natural rubber (Hevea brasiliensis). Mater. Sci. Eng. C 2021, 131, 112532. [Google Scholar] [CrossRef]
  53. Melli, V.; Juszczyk, M.; Sandrini, E.; Bolelli, G.; Bonferroni, B.; Lusvarghi, L.; Cigada, A.; Manfredini, T.; De Nardo, L. Tribological and mechanical performance evaluation of metal prosthesis components manufactured via metal injection molding. J. Mater. Sci.: Mater. Med. 2015, 26, 1–11. [Google Scholar] [CrossRef] [PubMed]
  54. Xu, P.; Pyczak, F.; Yan, M.; Kong, F.; Ebel, T. Impacts of yttrium on microstructure and tensile properties of biomedical β Ti-Nb-Zr fabricated by metal injection molding. Mater. Sci. Eng. A 2020, 792, 139816. [Google Scholar] [CrossRef]
  55. Suwanpreecha, C.; Alabort, E.; Tang, Y.T.; Panwisawas, C.; Reed, R.C.; Manonukul, A. A novel low-modulus titanium alloy for biomedical applications: A comparison between selective laser melting and metal injection moulding. Mater. Sci. Eng. A 2021, 812, 141081. [Google Scholar] [CrossRef]
  56. Aust, E.; Limberg, W.; Gerling, R.; Oger, B.; Ebel, T. Advanced TiAl6Nb7 bone screw implant fabricated by metal injection moulding. Adv. Eng. Mater. 2006, 8, 365–370. [Google Scholar] [CrossRef]
  57. Barbosa, A.P.C.; Bram, M.; Stöver, D.; Buchkremer, H.P. Realization of a titanium spinal implant with a gradient in porosity by 2-component-metal injection moulding. Adv. Eng. Mater. 2013, 15, 510–521. [Google Scholar] [CrossRef]
  58. Barbosa, A.P.C. Development of the 2-Component-Injection Moulding for Metal Powders; Forschungszentrum Jülich GmbH: Jülich, Germany, 2011. [Google Scholar]
  59. Shu, C.; He, H.; Fan, B.; Li, J.; Wang, T.; Li, D.; Li, Y.; He, H. Biocompatibility of vascular stents manufactured using metal inection molding in animal experiments. Trans. Nonferrous Met. Soc. China 2022, 32, 569–580. [Google Scholar] [CrossRef]
  60. Dehghan-Manshadi, A.; Yu, P.; Dargusch, M.; StJohn, D.; Qian, M. Metal injection moulding of surgical tools, biomaterials and medical devices: A review. Powder Technol. 2020, 364, 189–204. [Google Scholar] [CrossRef]
  61. Sidambe, A.T. Biocompatibility of advanced manufactured titanium implants—A review. Materials 2014, 7, 8168–8188. [Google Scholar] [CrossRef]
  62. Ezugwu, E.O.; Da Silva, R.B.; Sales, W.F.; Machado, A.R. Overview of the machining of titanium alloys. In Encyclopedia of Sustainable Technologies; Abraham, M.A., Ed.; Elsevier: Amsterdam, The Netherlands, 2017; pp. 487–506. [Google Scholar]
  63. Ronoh, K.; Mwema, F.; Dabees, S.; Sobola, D. Advances in sustainable grinding of different types of the titanium biomaterials for medical applications: A review. Biomed. Eng. Adv. 2022, 4, 100047. [Google Scholar] [CrossRef]
  64. Pede, D.; Li, M.; Virovac, L.; Poleske, T.; Balle, F.; Müller, C.; Mozaffari-Jovein, H. Microstructure and corrosion resistance of novel β-type titanium alloys manufactured by selective laser melting. J. Mater. Res. Technol. 2022, 19, 4598–4612. [Google Scholar] [CrossRef]
  65. Huang, Z.; Xiao, H.; Yu, J.; Zhang, H.; Huang, H.; Yu, K.; Zhou, R. Effects of different annealing cooling methods on the microstructure and properties of TA10 titanium alloys. J. Mater. Res. Technol. 2022, 18, 4859–4870. [Google Scholar] [CrossRef]
  66. Chen, K.; Fan, Q.; Yang, L.; Yao, J.; Xu, S.; Lei, W.; Gao, Y. Deciphering the microstructural evolution and adiabatic shearing behavior of the titanium alloy with stress-induced ω phase transformation during dynamic compression. Mater. Des. 2022, 221, 110939. [Google Scholar] [CrossRef]
  67. Yasin, M.S.; Soltani-Tehrani, A.; Shao, S.; Haghshenas, M.; Shamsaei, N. A comparative study on fatigue performance of various additively manufactured titanium alloys. Procedia Struct. Integr. 2022, 38, 519–525. [Google Scholar] [CrossRef]
  68. Moghadam, M.S.; Fayyaz, A.; Ardestani, M. Fabrication of titanium components by low-pressure powder injection moulding using hydride-dehydride titanium powder. Powder Technol. 2021, 377, 70–79. [Google Scholar] [CrossRef]
  69. Kim, S.; Jung, H.; Rim, H.J.; Lee, H.-S.; Lee, W. Fabrication of reinforced α+β titanium alloys by infiltration of Al into porous Ti-V compacts. J. Alloys Compd. 2018, 768, 775–781. [Google Scholar] [CrossRef]
  70. Dang, K.; Wang, K.; Chen, W.; Liu, G. Study on fast gas forming with in-die quenching for titanium alloys and the strengthening mechanisms of the components. J. Mater. Res. Technol. 2022, 18, 3916–3932. [Google Scholar] [CrossRef]
  71. Gupta, M.K.; Niesłony, P.; Sarikaya, M.; Korkmaz, M.E.; Kuntoğlu, M.; Królczyk, G.M.; Jamil, M. Tool wear patterns and their promoting mechanisms in hybrid cooling assisted machining of titanium Ti-3Al-2.5V/grade 9 alloy. Tribol. Int. 2022, 174, 107773. [Google Scholar] [CrossRef]
  72. Thavanayagam, G.; Swan, J.E. Optimizing hydride-dehydride Ti-6Al-4V feedstock composition for titanium powder injection moulding. Powder Technol. 2019, 355, 688–699. [Google Scholar] [CrossRef]
  73. De Freitas Daudt, N.; Bram, M.; Barbosa, A.P.C.; Laptev, A.M.; Alves, C. Manufacturing of highly porous titanium by metal injection molding in combination with plasma treatment. J. Mater. Process. Technol. 2017, 239, 202–209. [Google Scholar] [CrossRef]
  74. Nor, N.H.M.; Ismail, M.H.; Husain, H.; Saedon, J.B.; Newishy, M. Sintered strength of Ti-6Al-4V by metal injection molding (MIM) using palm stearin binder. J. Mech. Eng. 2019, 16, 135–147. [Google Scholar]
  75. Shbeh, M.M.; Goodall, R. Open pore titanium foams via metal injection molding of metal powder with a space holder. Met. Powder Rep. 2016, 71, 450–455. [Google Scholar] [CrossRef]
  76. Fattah-alhosseini, A.; Molaei, M.; Nouri, M.; Babaei, K. Review of the role of graphene and its derivatives in enhancing the performance of plasma electrolytic oxidation coatings on titanium and its alloys. Appl. Surf. Sci. Adv. 2021, 6, 100140. [Google Scholar] [CrossRef]
  77. Alves, S.A.; Bayon, R.; de Viteri, V.S.; Garcia, M.P.; Igartua, A.; Fernandes, M.H.; Rocha, L.A. Tribocorrosion behavior of calcium- and phosphorous-enriched titanium oxide films and study of osteoblast interactions for dental implants. J. Bio Tribo Corros. 2015, 1, 23. [Google Scholar] [CrossRef]
  78. Jiang, H. Enhancement of titanium alloy corrosion resistance via anodic oxidation treatment. Int. J. Electrochem. Sci. 2018, 13, 3888–3896. [Google Scholar] [CrossRef]
  79. Fazel, M.; Salimijazi, H.R.; Shamanian, M. Improvement of corrosion and tribocorrosion behavior of pure titanium by subzero anodic spark oxidation. ACS Appl. Mater. Interfaces 2018, 10, 15281–15287. [Google Scholar] [CrossRef]
  80. Venkateswarlu, K.; Rameshbabu, N.; Sreekanth, D.; Bose, A.C.; Muthupandi, V.; Babu, N.K.; Subramanian, S. Role of electrolyte additives on in-vitro electrochemical behavior of micro arc oxidized titania films on Cp Ti. Appl. Surf. Sci. 2012, 258, 6853–6863. [Google Scholar]
  81. Prasad, S.; Ehrensberger, M.; Gibson, M.P.; Kim, H.; Monaco, E.A. Biomaterial properties of titanium in dentistry. J. Oral Biosci. 2015, 57, 192–199. [Google Scholar] [CrossRef]
  82. Lautenschlager, E.P.; Monaghan, P. Titanium and titanium alloys as dental materials. Int. Dent. J. 1993, 43, 245–253. [Google Scholar]
  83. Zakaria, M.Y.; Sulong, A.B.; Muhamad, N.; Raza, M.R.; Ramli, M.I. Incorporation of wollastonite bioactive ceramic with titanium for medical applications: An overview. Mater. Sci. Eng. C 2019, 97, 884–895. [Google Scholar] [CrossRef]
  84. Oshida, Y. Bioscience and Bioengineering of Titanium Materials; Elsevier: Amsterdam, The Netherlands, 2007. [Google Scholar]
  85. Cabrera, N.; Mott, N.F. Theory of the oxidation of metals. Rep. Prog. Phys. 1949, 12, 163–184. [Google Scholar] [CrossRef]
  86. Chunxiang, C.; BaoMin, H.; Lichen, Z.; Shuangjin, L. Titanium alloy production technology, market prospects and industry development. Mater. Des. 2011, 32, 1684–1691. [Google Scholar]
  87. Ivasyshyn, O.M.; Aleksandrov, A.V. Status of the titanium production, research, and applications in the CIS. Mater. Sci. 2008, 44, 311–327. [Google Scholar] [CrossRef]
  88. Kearns, M. Titanium: Alive, well, and booming! Adv. Mater. Process. 2005, 163, 63–64. [Google Scholar]
  89. Brånemark, R.; Brånemark, P.-I.; Rydevik, B.; Myers, R.R. Osseointegration in skeletal reconstruction and rehabilitation. J. Rehabil. Res. Dev. 2001, 38, 175–181. [Google Scholar] [PubMed]
  90. Alipal, J.; Lee, T.C.; Koshy, P.; Abdullah, H.Z.; Idris, M.I. Evolution of anodised titanium for implant applications. Heliyon 2021, 7, e07408. [Google Scholar] [CrossRef] [PubMed]
  91. Kaluđerović, M.R.; Schreckenbach, J.P.; Graf, H.-L. Titanium dental implant surfaces obtained by anodic spark deposition—From the past to the future. Mater. Sci. Eng. C 2016, 69, 1429–1441. [Google Scholar] [CrossRef]
  92. Mohammadi, H.; Muhamad, N.; Sulong, A.B.; Ahmadipour, M. Recent advances on biofunctionalization of metallic substrate using ceramic coating: How far are we from clinically stable implant? J. Taiwan Inst. Chem. Eng. 2021, 118, 254–270. [Google Scholar] [CrossRef]
  93. Civantos, A.; Martínez-Campos, E.; Ramos, V.; Elvira, C.; Gallardo, A.; Abarrategi, A. Titanium coatings and surface modifications: Toward clinically useful bioactive implants. ACS Biomater. Sci. Eng. 2017, 3, 1245–1261. [Google Scholar] [CrossRef]
  94. Apostu, D.; Lucaciu, O.; Berce, C.; Lucaciu, D.; Cosma, D. Current methods of preventing aseptic loosening and improving osseointegration of titanium implants in cementless total hip arthroplasty: A review. J. Int. Med. Res. 2018, 46, 2104–2119. [Google Scholar] [CrossRef]
  95. Apostu, D.; Lucaciu, O.; Lucaciu, G.D.O.; Crisan, B.; Crisan, L.; Bacuit, M.; Onisor, F.; Baciut, G.; Câmpian, R.S.; Bran, S. Systemic drugs that influence titanium implant osseointegration. Drug Metab. Rev. 2017, 49, 92–104. [Google Scholar] [CrossRef]
  96. John, A.A.; Jaganathan, S.K.; Supriyanto, E.; Manikandan, A. Surface modification of titanium and its alloys for the enhancement of osseointegration in orthopaedics. Curr. Sci. 2016, 111, 1003–1015. [Google Scholar] [CrossRef]
  97. Lewallen, E.A.; Riester, S.M.; Bonin, C.A.; Kremers, H.M.; Dudakovic, A.; Kakar, S.; Cohen, R.C.; Westendorf, J.J.; Lewallen, D.G.; Wijnen, A.J.V. Biological strategies for improved osseointegration and osteoinduction of porous metal orthopedic implants. Tissue Eng. Part B Rev. 2015, 21, 218–230. [Google Scholar] [CrossRef] [PubMed]
  98. Sartoretto, S.C.; Alves, A.T.N.N.; Resende, R.F.B.; Calasans-Maia, J.; Granjeiro, J.M.; Calasans-Maia, M.D. Early osseointegration driven by the surface chemistry and wettability of dental implants. J. Appl. Oral Sci. 2015, 23, 279–287. [Google Scholar] [CrossRef] [PubMed]
  99. Manivasagam, G.; Dhinasekaran, D.; Rajamanickam, A. Biomedical implants: Corrosion and its prevention—A review. Recent Pat. Corros. Sci. 2010, 2, 40–54. [Google Scholar] [CrossRef]
  100. Elias, C.N.; Lima, J.H.C.; Valiev, R.; Meyers, M.A. Biomedical applications of titanium and its alloys. JOM 2008, 60, 46–49. [Google Scholar] [CrossRef]
  101. Paschalis, E.I.; Chodosh, J.; Spurr-Michaud, S.; Cruzat, A.; Tauber, A.; Behlau, I.; Gipson, I.; Dohlman, C.H. In vitro and in vivo assessment of titanium surface modification for coloring the backplate of the Boston keratoprosthesis. Investig. Ophthalmol. Vis. Sci. 2013, 54, 3863–3873. [Google Scholar] [CrossRef]
  102. Kärrholm, J.; Razaznejad, R. Fixation and bone remodeling around a low stiffness stem in revision surgery. Clin. Orthop. Relat. Res. 2008, 466, 380–388. [Google Scholar] [CrossRef]
  103. Niinimaki, T.; Jalovaara, P. Bone loss from the proximal femur after arthroplasty with an isoelastic femoral stem: BMD measurements in 25 patients after 9 years. Acta Orthop. Scand. 1995, 66, 347–351. [Google Scholar] [CrossRef]
  104. Spoerke, E.D.; Murray, N.G.; Li, H.; Brinson, L.C.; Dunand, D.C.; Stupp, S.I. A bioactive titanium foam scaffold for bone repair. Acta Biomater. 2005, 1, 523–533. [Google Scholar] [CrossRef]
  105. Niinomi, M. Mechanical biocompatibilities of titanium alloys for biomedical applications. J. Mech. Behav. Biomed. Mater. 2008, 1, 30–42. [Google Scholar] [CrossRef]
  106. Li, F.; Li, J.; Xu, G.; Liu, G.; Kou, H.; Zhou, L. Fabrication, pore structure and compressive behavior of anisotropic porous titanium for human trabecular bone implant applications. J. Mech. Behav. Biomed. Mater. 2015, 46, 104–114. [Google Scholar] [CrossRef]
  107. Johari, N.; Fathi, M.H.; Golozar, M.A. Fabrication, characterization and evaluation of the mechanical properties of poly (ε-caprolactone)/nano-fluoridated hydroxyapatite scaffold for bone tissue engineering. Compos. Part B 2012, 43, 1671–1675. [Google Scholar] [CrossRef]
  108. Han, C.; Li, Y.; Wu, X.; Ren, S.; San, X.; Zhu, X. Ti/SiO2 composite fabricated by powder metallurgy for orthopedic implant. Mater. Des. 2013, 49, 76–80. [Google Scholar] [CrossRef]
  109. Chen, Q.; Thouas, G.A. Metallic implant biomaterials. Mater. Sci. Eng. R Rep. 2015, 87, 1–57. [Google Scholar] [CrossRef]
  110. Katti, K.S. Biomaterials in total joint replacement. Colloids Surf. B Biointerfaces 2004, 39, 133–142. [Google Scholar] [CrossRef] [PubMed]
  111. Leyens, C.; Peters, M. Titanium and Titanium Alloys: Fundamentals and Applications; John Wiley & Sons: Weinheim, Germany, 2006. [Google Scholar]
  112. Lampman, S. Titanium and its alloys for biomedical implants. In ASM Handbook Materials for Medical Devices; Narayan, R., Ed.; ASM International: Russell Township, OH, USA, 2012; pp. 225–238. [Google Scholar]
  113. Kafkas, F.; Ebel, T. Metallurgical and mechanical properties of Ti-24Nb-4Zr-8Sn alloy fabricated by metal injection molding. J. Alloys Compd. 2014, 617, 359–366. [Google Scholar] [CrossRef]
  114. Vishnu, J.; Sankar, M.; Rack, H.J.; Rao, N.; Singh, A.K.; Manivasagam, G. Effect of phase transformations during aging on tensile strength and ductility of metastable beta titanium alloy Ti-35Nb-7Zr-5Ta-0.35O for orthopedic applications. Mater. Sci. Eng. A 2020, 779, 139127. [Google Scholar] [CrossRef]
  115. Niinomi, M. Mechanical properties of biomedical titanium alloys. Mater. Sci. Eng. A 1998, 243, 231–236. [Google Scholar] [CrossRef]
  116. Geetha, M.; Singh, A.K.; Asokamani, R.; Gogia, A.K. Ti based biomaterials, the ultimate choice for orthopaedic implants—A review. Prog. Mater. Sci. 2009, 54, 397–425. [Google Scholar] [CrossRef]
  117. Hynowska, A.; Pellicer, E.; Fornell, J.; González, S.; van Steenberge, N.; Suriñach, S.; Gebert, A.; Calin, M.; Eckert, J.; Baró, M.D.; et al. Nanostructured β-phase Ti-31.0Fe-9.0Sn and sub-μm structured Ti-39.3Nb-13.3Zr-10.7Ta alloys for biomedical applications: Microstructure benefits on the mechanical and corrosion performances. Mater. Sci. Eng. C 2012, 32, 2418–2425. [Google Scholar] [CrossRef]
  118. Rack, H.J.; Qazi, J.I. Titanium alloys for biomedical applications. Mater. Sci. Eng. C 2006, 26, 1269–1277. [Google Scholar] [CrossRef]
  119. Bauer, S.; Schmuki, P.; von der Mark, K.; Park, J. Engineering biocompatible implant surfaces: Part I: Materials and surfaces. Prog. Mater. Sci. 2013, 58, 261–326. [Google Scholar] [CrossRef]
  120. Deligianni, D.D.; Katsala, N.; Ladas, S.; Sotiropoulou, D.; Amedee, J.; Missirlis, Y.F. Effect of surface roughness of the titanium alloy Ti-6Al-4V on human bone marrow cell response and on protein adsorption. Biomaterials 2001, 22, 1241–1251. [Google Scholar] [CrossRef] [PubMed]
  121. Wan, Y.; Zhao, Z.; Yu, M.; Ji, Z.; Wang, T.; Cai, Y.; Liu, C.; Liu, Z. Osteogenic and antibacterial ability of micro-nano structures coated with ZnO on Ti-6Al-4V implant fabricated by two-step laser processing. J. Mater. Sci. Technol. 2022, 131, 240–252. [Google Scholar] [CrossRef]
  122. Lantang, Y.S.F.; Kobayashi, E.; Sato, T. Fabrication of spark plasma sintered porous Ti-6Al-4V using PCA-added space holder method for bone implant materials. Mater. Today Proc. 2022, 66, 2693–2701. [Google Scholar] [CrossRef]
  123. Dewidar, M.M.; Lim, J.K. Properties of solid core and porous surface Ti-6Al-4V implants manufactured by powder metallurgy. J. Alloys Compd. 2008, 454, 442–446. [Google Scholar] [CrossRef]
  124. Vazirgiantzikis, I.; George, S.L.; Pichon, L. Surface characterisation and silver release from Ti-6Al-4V and anodic TiO2 after surface modification by ion implantation. Surf. Coat. Technol. 2022, 433, 128115. [Google Scholar] [CrossRef]
  125. Lim, B.-S.; Cho, H.-R.; Choe, H.-C. Corrosion behaviors of macro/micro/nano-scale surface modification on Ti-6Al-4V alloy for bio-implant. Thin Solid Films 2022, 754, 139314. [Google Scholar] [CrossRef]
  126. Rahmouni, K.; Besnard, A.; Oulmi, K.; Nouveau, C.; Hidoussi, A.; Aissani, L.; Zaabat, M. In vitro corrosion response of CoCrMo and Ti-6Al-4V orthopedic implants with Zr columnar thin films. Surf. Coat. Technol. 2022, 436, 128310. [Google Scholar] [CrossRef]
  127. Zhao, B.; Wang, H.; Qiao, N.; Wang, C.; Hu, M. Corrosion resistance characteristics of a Ti-6Al-4V alloy scaffold that is fabricated by electron beam melting and selective laser melting for implantation in vivo. Mater. Sci. Eng. C 2017, 70, 832–841. [Google Scholar] [CrossRef]
  128. Nagentrau, M.; Tobi, A.L.M.; Jamian, S.; Otsuka, Y.; Hussin, R. Delamination-fretting wear failure evaluation at HAp-Ti-6Al–4V interface of uncemented artificial hip implant. J. Mech. Behav. Biomed. Mater. 2021, 122, 104657. [Google Scholar] [CrossRef]
  129. Huang, H.-H.; Shiau, D.-K.; Chen, C.-S.; Chang, J.-H.; Wang, S.; Pan, H.; Wu, M.-F. Nitrogen plasma immersion ion implantation treatment to enhance corrosion resistance, bone cell growth, and antibacterial adhesion of Ti-6Al-4V alloy in dental applications. Sur. Coat. Technol. 2019, 365, 179–188. [Google Scholar] [CrossRef]
  130. Ren, B.; Wan, Y.; Liu, C.; Wang, H.; Yu, M.; Zhang, X.; Huang, Y. Improved osseointegration of 3D printed Ti-6Al-4V implant with a hierarchical micro/nano surface topography: An in vitro and in vivo study. Mater. Sci. Eng. C 2021, 118, 111505. [Google Scholar] [CrossRef] [PubMed]
  131. Du, Z.; Xiao, S.; Liu, J.; Lv, S.; Xu, L.; Kong, F.; Chen, Y. Hot deformation behavior of Ti-3.5Al-5Mo-6V-3Cr-2Sn-0.5Fe Alloy in α + β field. Metals 2015, 5, 216–227. [Google Scholar] [CrossRef]
  132. Akahori, T.; Niinomi, M.; Fukui, H.; Ogawa, M.; Toda, H. Improvement in fatigue characteristics of newly developed beta type titanium alloy for biomedical applications by thermo-mechanical treatments. Mater. Sci. Eng. C 2005, 25, 248–254. [Google Scholar] [CrossRef]
  133. Manam, N.S.; Harun, W.S.W.; Shri, D.N.A.; Ghani, S.A.C.; Kurniawan, T.; Ismail, M.H.; Ibrahim, M.H.I. Study of corrosion in biocompatible metals for implants: A review. J. Alloys Compd. 2017, 701, 698–715. [Google Scholar] [CrossRef]
  134. Liu, H.; Niinomi, M.; Nakai, M.; Obara, S.; Fujii, H. Improved fatigue properties with maintaining low young’s modulus achieved in biomedical beta-type titanium alloy by oxygen addition. Mater. Sci. Eng. A 2017, 704, 10–17. [Google Scholar] [CrossRef]
  135. Afzali, P.; Ghomashchi, R.; Oskouei, R.H. On the corrosion behaviour of low modulus titanium alloys for medical implant applications: A review. Metals 2019, 9, 878. [Google Scholar] [CrossRef]
  136. Yilmazer, H.; Niinom, M.; Nakai, M.; Hieda, J.; Todaka, Y.; Akahori, T.; Miyazaki, T. Heterogeneous structure and mechanical hardness of biomedical β-type Ti-29Nb-13Ta-4.6Zr subjected to high-pressure torsion. J. Mech. Behav. Biomed. Mater. 2012, 10, 235–245. [Google Scholar] [CrossRef]
  137. Takematsu, E.; Cho, K.; Hieda, J.; Nakai, M.; Katsumata, K.; Okada, K.; Niinomi, M.; Matsushita, N. Adhesive strength of bioactive oxide layers fabricated on TNTZ alloy by three different alkali-solution treatments. J. Mech. Behav. Biomed. Mater. 2016, 61, 174–181. [Google Scholar] [CrossRef]
  138. Raducanu, D.; Vasilescu, E.; Cojocaru, V.D.; Cinca, I.; Drob, P.; Vasilescu, C.; Drob, S.I. Mechanical and corrosion resistance of a new nanostructured Ti-Zr-Ta-Nb alloy. J. Mech. Behav. Biomed. Mater. 2011, 4, 1421–1430. [Google Scholar] [CrossRef] [PubMed]
  139. Niinomi, M.; Hattori, T.; Morikawa, K.; Kasuga, T.; Suzuki, A.; Fukui, H.; Niwa, S. Development of low rigidity β-type titanium alloy for biomedical applications. Mater. Trans. 2002, 43, 2970–2977. [Google Scholar] [CrossRef]
  140. Niinomi, M. Fatigue performance and cyto-toxicity of low rigidity titanium alloy, Ti-29Nb-13Ta-4.6Zr. Biomaterials 2003, 24, 2673–2683. [Google Scholar] [CrossRef] [PubMed]
  141. Milošev, I.; Žerjav, G.; Moreno, J.M.C.; Popa, M. Electrochemical properties, chemical composition and thickness of passive film formed on novel Ti-20Nb-10Zr-5Ta alloy. Electrochim. Acta 2013, 99, 176–189. [Google Scholar] [CrossRef]
  142. German, R.M. Powder Injection Moulding; Metal Powder Industries Federation: Princeton, NJ, USA, 1990. [Google Scholar]
  143. Emeka, U.B.; Sulong, A.B.; Muhamad, N.; Sajuri, Z.; Salleh, F. Two component injection molding of bi-material of stainless steel and yttria stabilized zirconia—Green Part. J. Kejuruter. 2017, 29, 49–55. [Google Scholar] [CrossRef]
  144. Fayyaz, A.; Muhamad, N.; Sulong, A.B. Microstructure and physical and mechanical properties of micro cemented carbide injection moulded components. Powder Technol. 2018, 326, 151–158. [Google Scholar] [CrossRef]
  145. Jabir, S.M.; Noorsyakirah, A.; Afian, O.M.; Nurazilah, M.Z.; Aswad, M.A.; Afiq, N.H.M.; Mazlan, M. Analysis of the rheological behavior of copper metal injection molding (MIM) feedstock. Procedia Chem. 2016, 19, 148–152. [Google Scholar] [CrossRef]
  146. Antusch, S.; Armstrong, D.E.J.; Britton, T.B.; Commin, L.; Gibson, J.S.K.-L.; Greuner, H.; Hoffmann, J.; Knabl, W.; Pintsuk, G.; Rieth, M.; et al. Mechanical and microstructural investigations of tungsten and doped tungsten materials produced via powder injection molding. Nucl. Mater. Energy 2015, 3–4, 22–31. [Google Scholar] [CrossRef]
  147. Thavanayagam, G.; Pickering, K.L.; Swan, J.E.; Cao, P. Analysis of rheological behaviour of titanium feedstocks formulated with a water-soluble binder system for powder injection moulding. Powder Technol. 2015, 269, 227–232. [Google Scholar] [CrossRef]
  148. You, W.-K.; Choi, J.-P.; Yoon, S.-M.; Lee, J.-S. Low temperature powder injection molding of iron micro-nano powder mixture. Powder Technol. 2012, 228, 199–205. [Google Scholar] [CrossRef]
  149. Mukund, B.N.; Hausnerova, B.; Shivashankar, T.S. Development of 17-4PH stainless steel bimodal powder injection molding feedstock with the help of interparticle spacing/lubricating liquid concept. Powder Technol. 2015, 283, 24–31. [Google Scholar] [CrossRef]
  150. Wang, J.; Edirisinghe, M.J. Ceramic injection molding. In Reference Module in Materials Science and Materials Engineering; Elsevier: Amsterdam, The Netherlands, 2016; pp. 1–22. [Google Scholar]
  151. Heaney, D.F. Handbook of Metal Injection Molding; Woodhead Publishing: Cambridge, UK, 2012. [Google Scholar]
  152. Piotter, V.; Bauer, W.; Knitter, R.; Mueller, M.; Mueller, T.; Plewa, K. Powder injection moulding of metallic and ceramic micro parts. Microsyst. Technol. 2011, 17, 251–263. [Google Scholar] [CrossRef]
  153. Tuncer, N.; Bram, M.; Laptev, A.; Beck, T.; Moser, A.; Buchkremer, H.P. Study of metal injection molding of highly porous titanium by physical modeling and direct experiments. J. Mater. Process. Technol. 2014, 214, 1352–1360. [Google Scholar] [CrossRef]
  154. Shbeh, M.M.; Yerokhin, A.; Goodall, R. Microporous titanium through metal injection moulding of coarse powder and surface modification by plasma oxidation. Appl. Sci. 2017, 7, 105. [Google Scholar] [CrossRef]
  155. Li, J.; Liu, B.; Zhou, Y.; Chen, Z.; Jiang, L.; Yuan, W.; Liang, L. Fabrication of a Ti porous microneedle array by metal injection molding for transdermal drug delivery. PLoS ONE 2017, 12, e0172043. [Google Scholar] [CrossRef]
  156. Dehghan-Manshadi, A.; Chen, Y.; Shi, Z.; Bermingham, M.; StJohn, D.; Dargusch, M.; Qian, M. Porous titanium scaffolds fabricated by metal injection moulding for biomedical applications. Materials 2018, 11, 1573. [Google Scholar] [CrossRef]
  157. Hayat, M.D.; Goswami, A.; Matthews, S.; Li, T.; Yuan, X.; Cao, P. Modification of PEG/PMMA binder by PVP for titanium metal injection moulding. Powder Technol. 2017, 315, 243–249. [Google Scholar] [CrossRef]
  158. Hayat, M.D.; Jadhav, P.P.; Zhang, H.; Ray, S.; Cao, P. Improving titanium injection moulding feedstock based on PEG/PPC based binder system. Powder Technol. 2018, 330, 304–309. [Google Scholar] [CrossRef]
  159. Hayat, M.D.; Wen, G.; Zulkifli, M.F.; Cao, P. Effect of PEG molecular weight on rheological properties of Ti-MIM feedstocks and water debinding behaviour. Powder Technol. 2015, 270, 296–301. [Google Scholar] [CrossRef]
  160. Bootchai, S.; Taweejun, N.; Manonukul, A.; Kanchanomai, C. Metal injection molded titanium: Mechanical properties of debinded powder and sintered metal. J. Mater. Eng. Perform. 2020, 29, 4559–4568. [Google Scholar] [CrossRef]
  161. Zheng, J.-p.; Chen, L.-j.; Chen, D.-y.; Shao, C.-s.; Yi, M.-f.; Zhang, B. Effects of pore size and porosity of surface-modified porous titanium implants on bone tissue ingrowth. Trans. Nonferrous Met. Soc. China 2019, 29, 2534–2545. [Google Scholar] [CrossRef]
  162. Chen, G.; Cao, P.; Wen, G.; Edmonds, N.; Li, Y. Using an agar-based binder to produce porous NiTi alloys by metal injection moulding. Intermetallics 2013, 37, 92–99. [Google Scholar] [CrossRef]
  163. Xu, W.; Lu, X.; Wang, L.N.; Shi, Z.M.; Lv, S.M.; Qian, M.; Qu, X.H. Mechanical properties, in vitro corrosion resistance and biocompatibility of metal injection molded Ti-12Mo alloy for dental applications. J. Mech. Behav. Biomed. Mater. 2018, 88, 534–547. [Google Scholar] [CrossRef] [PubMed]
  164. Cho, K.; Niinomi, M.; Nakai, M.; Liu, H.; Santos, P.F.; Itoh, Y.; Ikeda, M.; Abdel-Hady Gepreel, M.; Narushima, T. Improvement in mechanical strength of low-cost β-type Ti-Mn alloys fabricated by metal injection molding through cold rolling. J. Alloys Compd. 2016, 664, 272–283. [Google Scholar] [CrossRef]
  165. Santos, P.F.; Niinomi, M.; Liu, H.; Cho, K.; Nakai, M.; Itoh, Y.; Narushima, T.; Ikeda, M. Fabrication of low-cost beta-type Ti-Mn alloys for biomedical applications by metal injection molding process and their mechanical properties. J. Mech. Behav. Biomed. Mater. 2016, 59, 497–507. [Google Scholar] [CrossRef]
  166. Zhao, D.; Chang, K.; Ebel, T.; Qian, M.; Willumeit, R.; Yan, M.; Pyczak, F. Microstructure and mechanical behavior of metal injection molded Ti-Nb binary alloys as biomedical material. J. Mech. Behav. Biomed. Mater. 2013, 28, 171–182. [Google Scholar] [CrossRef]
  167. Zhao, D.; Chang, K.; Ebel, T.; Nie, H.; Willumeit, R.; Pyczak, F. Sintering behavior and mechanical properties of a metal injection molded Ti-Nb binary alloy as biomaterial. J. Alloys Compd. 2015, 640, 393–400. [Google Scholar] [CrossRef]
  168. Yılmaz, E.; Gökçe, A.; Findik, F.; Gulsoy, H. Metallurgical properties and biomimetic HA deposition performance of Ti-Nb PIM alloys. J. Alloys Compd. 2018, 746, 301–313. [Google Scholar] [CrossRef]
  169. Bidaux, J.-E.; Pasquier, R.; Rodriguez-Arbaizar, M.; Girard, H.; Carren~o-Morelli, E. Low elastic modulus Ti-17Nb processed by powder injection moulding and postsintering heat treatments. Powder Metall. 2014, 57, 320–323. [Google Scholar] [CrossRef]
  170. Zhao, D.-p.; Chen, Y.-k.; Chang, K.-k.; Ebel, T.; Luthrigner-Feyerabend, B.J.C.; Willumeit-Römer, R.; Pyczak, F. Surface topography and cytocompatibility of metal injection molded Ti-22Nb alloy as biomaterial. Trans. Nonferrous Met. Soc. China 2018, 28, 1342–1350. [Google Scholar] [CrossRef]
  171. Holm, M.; Ebel, T.; Dahms, M. Investigations on Ti-6Al-4V with gadolinium addition fabricated by metal injection moulding. Mater. Des. 2013, 51, 943–948. [Google Scholar] [CrossRef]
  172. Bian, T.; Ding, C.; Yao, X.; Wang, J.; Mo, W.; Wang, Z.; Yu, P.; Ye, S. Mechanisms of heat treatment and ductility improvement of high-oxygen Ti-6Al-4V alloy fabricated by metal injection molding. Mater. Sci. Eng. A 2022, 840, 142924. [Google Scholar] [CrossRef]
  173. Lin, D.; Sanetrnik, D.; Cho, H.; Chung, S.T.; Kwon, Y.S.; Kate, K.H.; Hausnerova, B.; Atre, S.V.; Park, S.J. Rheological and thermal debinding properties of blended elemental Ti-6Al-4V powder injection molding feedstock. Powder Technol. 2017, 311, 357–363. [Google Scholar] [CrossRef]
  174. Nor, N.H.M.; Muhamad, N.; Ihsan, A.K.A.M.; Jamaludin, K.R. Sintering parameter optimization of Ti-6Al-4V metal injection molding for highest strength using palm stearin binder. Procedia Eng. 2013, 68, 359–364. [Google Scholar] [CrossRef]
  175. Zakaria, M.Y.; Ramli, M.I.; Sulong, A.B.; Muhamad, N.; Ismail, M.H. Application of sodium chloride as space holder for powder injection molding of alloy titanium-hydroxyapatite composites. J. Mater. Res. Technol. 2021, 12, 478–486. [Google Scholar] [CrossRef]
  176. Arifin, A.; Sulong, A.B.; Muhamad, N.; Syarif, J.; Ramli, M.I. Powder injection molding of HA/Ti6Al4V composite using palm stearin as based binder for implant material. Mater. Des. 2015, 65, 1028–1034. [Google Scholar] [CrossRef]
  177. Ramli, M.I.; Sulong, A.B.; Muhamad, N.; Muchtar, A.; Zakaria, M.Y. Effect of sintering on the microstructure and mechanical properties of alloy titanium-wollastonite composite fabricated by powder injection moulding process. Ceram. Int. 2019, 45, 11648–11653. [Google Scholar] [CrossRef]
  178. Yılmaz, E.; Gökçe, A.; Findik, F.; Gülsoy, H.Ö. Characterization of biomedical Ti-16Nb-(0-4)Sn alloys produced by powder injection molding. Vacuum 2017, 142, 164–174. [Google Scholar] [CrossRef]
  179. Dehghan-Manshadi, A.; Bermingham, M.J.; Dargusch, M.S.; StJohn, D.H.; Qian, M. Metal injection moulding of titanium and titanium alloys: Challenges and recent development. Powder Technol. 2017, 319, 289–301. [Google Scholar] [CrossRef]
  180. Virdhian, S.; Osada, T.; Kang, H.G.; Tsumori, F.; Miura, H. Evaluation and analysis of distortion of complex shaped Ti-6Al-4V compacts by metal injection molding process. Key Eng. Mater. 2012, 520, 187–194. [Google Scholar] [CrossRef]
  181. Ferri, O.M.; Ebel, T.; Bormann, R. The influence of a small boron addition on the microstructure and mechanical properties of Ti-6Al-4V fabricated by metal injection moulding. Adv. Eng. Mater. 2011, 13, 436–447. [Google Scholar] [CrossRef]
  182. Ebel, T.; Blawert, C.; Willumeit, R.; Luthringer, B.J.C.; Ferri, O.M.; Feyerabend, F. Ti-6Al-4V–0.5B--a modified alloy for implants produced by metal injection molding. Adv. Eng. Mater. 2011, 13, B440–B453. [Google Scholar] [CrossRef]
  183. Uematsu, T.; Itoh, Y.; Sato, K.; Miura, H. Effects of substrate for sintering on the mechanical properties of injection molded Ti-6Al-4V alloy. J. Jpn. Soc. Powder Metall. 2006, 53, 755–759. [Google Scholar] [CrossRef]
  184. Tingskong, T.; Larouche, F.; Lefebvre, L.-P. New titanium alloy feedstock for high performance metal injection moulding parts. Key Eng. Mater. 2016, 704, 118–121. [Google Scholar] [CrossRef]
  185. Pelletier, R.; Lefebvre, L.-P.; Baril, E. A MIM route for producing Ti6Al4V-TiC composites. Key Eng. Mater. 2016, 704, 139–147. [Google Scholar] [CrossRef]
  186. Obasi, G.C.; Ferri, O.M.; Ebel, T.; Bormann, R. Influence of processing parameters on mechanical properties of Ti-6Al-4V alloy fabricated by MIM. Mater. Sci. Eng. A 2010, 527, 3929–3935. [Google Scholar] [CrossRef]
  187. Attia, U.M.; Alcock, J.R. A review of micro-powder injection moulding as a microfabrication technique. J. Micromech. Microeng. 2011, 21, 043001. [Google Scholar] [CrossRef]
  188. Baek, E.-R.; Supriadi, S.; Choi, C.-J.; Lee, B.-T.; Lee, J.-W. Effect of particle size in feedstock properties in micro powder injection molding. Mater. Sci. Forum 2007, 534–536, 349–352. [Google Scholar] [CrossRef]
  189. Kusaka, K.; Kohno, T.; Kondo, T.; Horata, A. Tensile behavior of sintered titanium by MIM process. J. Jpn. Soc. Powder Metall. 1995, 42, 383–387. [Google Scholar] [CrossRef]
  190. Carreño-Morelli, E.; Rodríguez-Arbaizar, M.; Amherd, A.; Bidaux, J.-E. Porous titanium processed by powder injection moulding of titanium hydride and space holders. Powder Metall. 2014, 57, 93–96. [Google Scholar] [CrossRef]
  191. Carreño-Morelli, E.; Amherd, A.; Rodriguez-Arbaizar, M.; Zufferey, D.; Várez, A.; Bidaux, J.-E. Porous titanium by powder injection moulding of titanium hydride and PMMA space holders. Eur. Cells Mater. 2013, 26, 16. [Google Scholar]
  192. Carreño-Morelli, E.; Bidaux, J.-E.; Rodríguez-Arbaizar, M.; Girard, H.; Hamdan, H. Production of titanium grade 4 components by powder injection moulding of titanium hydride. Powder Metall. 2014, 57, 89–92. [Google Scholar] [CrossRef]
  193. German, R.M. Progress in titanium metal powder injection molding. Materials 2013, 6, 3641–3662. [Google Scholar] [CrossRef] [PubMed]
  194. Tosello, G. Micro Injection Molding; Hanser Verlag GmbH & Co. KG: Munich, Germany, 2018. [Google Scholar]
  195. Mahmud, N.N.; Azam, F.A.A.; Ramli, M.I.; Foudzi, F.M.; Ameyama, K.; Sulong, A.B. Rheological properties of irregular-shaped titanium-hydroxyapatite bimodal powder composite moulded by powder injection moulding. J. Mater. Res. Technol. 2021, 11, 2255–2264. [Google Scholar] [CrossRef]
  196. Iriany, M.N.; Sahari, J.; Arifin, A.K. Stability of metal injection molding feedstock containing palm oil. J. Inst. Mater. Malaysia 2001, 2, 71–83. [Google Scholar]
  197. Basir, A.; Sulong, A.B.; Jamadon, N.H.; Muhamad, N. Sintering behavior of bi-material micro-component of 17-4PH stainless steel and yttria-stabilized zirconia produced by two-component micro-powder injection molding process. Materials 2022, 15, 2059. [Google Scholar] [CrossRef]
  198. Arifin, A.; Sulong, A.B.; Gunawan; Mohruni, A.S.; Yani, I. Palm stearin for alternative binder for MIM: A review. J. Ocean. Mech. Aerosp. 2014, 7, 18–23. [Google Scholar]
  199. German, R.M.; Bose, A. Injection Molding of Metals and Ceramics; Metal Powder Industries Federation: Princeton, NJ, USA, 1997. [Google Scholar]
  200. Fayyaz, A.; Muhamad, N.; Sulong, A.B.; Yunn, H.S.; Amin, S.Y.M.; Rajabi, J. Micro-powder injection molding of cemented tungsten carbide: Feedstock preparation and properties. Ceram. Int. 2015, 41, 3605–3612. [Google Scholar] [CrossRef]
  201. Fayyaz, A.; Muhamad, N.; Sulong, A.B.; Rajabi, J.; Wong, Y.N. Fabrication of cemented tungsten carbide components by micro-powder injection moulding. J. Mater. Process. Technol. 2014, 214, 1436–1444. [Google Scholar] [CrossRef]
  202. Sotomayor, M.E.; Várez, A.; Levenfeld, B. Influence of powder particle size distribution on rheological properties of 316L powder injection moulding feedstocks. Powder Technol. 2010, 200, 30–36. [Google Scholar] [CrossRef]
  203. Li, Y.; Li, L.; Khalil, K.A. Effect of powder loading on metal injection molding stainless steels. J. Mater. Process. Technol. 2007, 183, 432–439. [Google Scholar] [CrossRef]
  204. Zakaria, H.; Muhamad, N.; Sulong, A.B.; Ibrahim, M.H.I.; Foudzi, F. Moldability characteristics of 3 mol% yttria stabilized zirconia feedstock for micro-powder injection molding process. Sains Malays. 2014, 43, 129–136. [Google Scholar]
  205. Aggarwal, G.; Park, S.J.; Smid, I. Development of niobium powder injection molding: Part I. Feedstock and injection molding. Int. J. Refract. Met. Hard Mater. 2006, 24, 253–262. [Google Scholar] [CrossRef]
  206. Kong, X.; Barriere, T.; Gelin, J.C. Determination of critical and optimal powder loadings for 316L fine stainless steel feedstocks for micro-powder injection molding. J. Mater. Process. Technol. 2012, 212, 2173–2182. [Google Scholar] [CrossRef]
  207. Mutsuddy, B.C.; Ford, R.G. Ceramic Injection Molding; Chapman & Hall: London, UK, 1995. [Google Scholar]
  208. Reddy, J.J.; Vijayakumar, M.; Mohan, T.R.R.; Ramakrishnan, P. Loading of solids in a liquid medium: Determination of CBVC by torque rheometry. J. Eur. Ceram. Soc. 1996, 16, 567–574. [Google Scholar] [CrossRef]
  209. Zakaria, M.Y.; Sulong, A.B.; Muhamad, N.; Ramli, M.I. Rheological properties of titanium-hydroxyapatite with powder space holder composite feedstock for powder injection moulding. Int. J. Adv. Manuf. Technol. 2019, 102, 2591–2599. [Google Scholar] [CrossRef]
  210. Huang, B.; Liang, S.; Qu, X. The rheology of metal injection molding. J. Mater. Process. Technol. 2003, 137, 132–137. [Google Scholar] [CrossRef]
  211. Agote, I.; Odriozola, A.; Gutierrez, M.; Santamaría, A.; Quintanilla, J.; Coupelle, P.; Soares, J. Rheological study of waste porcelain feedstocks for injection moulding. J. Eur. Ceram. Soc. 2001, 21, 2843–2853. [Google Scholar] [CrossRef]
  212. Basir, A.; Sulong, A.B.; Jamadon, N.H.; Muhamad, N. Bi-material micro-part of stainless steel and zirconia by two-component micro-powder injection molding: Rheological properties and solvent debinding behavior. Metals 2020, 10, 595. [Google Scholar] [CrossRef]
  213. Yang, W.-W.; Yang, K.-Y.; Wang, M.-C.; Hon, M.-H. Solvent debinding mechanism for alumina injection molded compacts with water-soluble binders. Ceram. Int. 2003, 29, 745–756. [Google Scholar] [CrossRef]
  214. Lin, S.T.; German, R.M. Extraction debinding of injection-molded parts by condensed solvent. Int. J. Powder Metall. 1989, 21, 19–24. [Google Scholar]
  215. Aggarwal, G.; Smid, I.; Park, S.J.; German, R.M. Development of niobium powder injection molding. Part II: Debinding and sintering. Int. J. Refract. Met. Hard Mater. 2007, 25, 226–236. [Google Scholar] [CrossRef]
  216. Enneti, R.K.; Park, S.J.; German, R.M.; Atre, S.V. Review: Thermal debinding process in particulate materials processing. Mater. Manuf. Process. 2012, 27, 103–118. [Google Scholar] [CrossRef]
  217. Aslam, M.; Ahmad, F.; Yusoff, P.S.M.B.M.; Altaf, K.; Omar, M.A.; German, R.M. Powder injection molding of biocompatible stainless steel biodevices. Powder Technol. 2016, 295, 84–95. [Google Scholar] [CrossRef]
  218. Lame, O.; Bellet, D.; Michiel, M.D.; Bouvard, D. In situ microtomography investigation of metal powder compacts during sintering. Nucl. Instrum. Methods Phys. Res. Sect. B 2003, 200, 287–294. [Google Scholar] [CrossRef]
  219. Li, K. Ceramic Membranes for Separation and Reaction; John Wiley & Sons: Weinheim, Germany, 2007. [Google Scholar]
  220. Luo, S.D.; Song, T.; Lu, S.L.; Liu, B.; Tian, J.; Qian, M. High oxygen-content titanium and titanium alloys made from powder. J. Alloys Compd. 2020, 836, 155526. [Google Scholar] [CrossRef]
  221. Zhao, D.; Ebel, T.; Yan, M.; Qian, M. Trace carbon in biomedical beta-titanium alloys: Recent progress. JOM 2015, 67, 2236–2243. [Google Scholar] [CrossRef]
  222. Otsuki, B.; Takemoto, M.; Fujibayashi, S.; Neo, M.; Kokubo, T.; Nakamura, T. Pore throat size and connectivity determine bone and tissue ingrowth into porous implants: Three-dimensional micro-CT based structural analyses of porous bioactive titanium implants. Biomaterials 2006, 27, 5892–5900. [Google Scholar] [CrossRef]
  223. Itälä, A.I.; Ylänen, H.O.; Ekholm, C.; Karlsson, K.H.; Aro, H.T. Pore diameter of more than 100 µm is not requisite for bone ingrowth in rabbits. J. Biomed. Mater. Res. 2001, 58, 679–683. [Google Scholar] [CrossRef]
  224. Ye, H.; Liu, X.Y.; Hong, H. Characterization of sintered titanium/hydroxyapatite biocomposite using FTIR spectroscopy. J. Mater. Sci. Mater. Med. 2009, 20, 843–850. [Google Scholar] [CrossRef]
  225. Chu, C.; Lin, P.; Dong, Y.; Xue, X.; Zhu, J.; Yin, Z. Fabrication and characterization of hydroxyapatite reinforced with 20 vol% Ti particles for use as hard tissue replacement. J. Mater. Sci. Mater. Med. 2002, 13, 985–992. [Google Scholar] [CrossRef] [PubMed]
Figure 1. MIM-processed medical tools and equipment, reused with permission from Elsevier [60].
Figure 1. MIM-processed medical tools and equipment, reused with permission from Elsevier [60].
Materials 16 03991 g001
Figure 2. Bone screw design for MIM processing and the placement of a typical implant in the second cervical vertebra of the neck, reused with permission from John Wiley and Sons [56].
Figure 2. Bone screw design for MIM processing and the placement of a typical implant in the second cervical vertebra of the neck, reused with permission from John Wiley and Sons [56].
Materials 16 03991 g002
Figure 3. (a) Spinal implant of titanium produced employing the two-component-MIM process; (b) positioning of implant immediately after surgical procedure; and (c) cross-sectional microstructure exhibiting adequate joining between the dense parts and porous, reused with permission from Elsevier and John Wiley and Sons [57,60].
Figure 3. (a) Spinal implant of titanium produced employing the two-component-MIM process; (b) positioning of implant immediately after surgical procedure; and (c) cross-sectional microstructure exhibiting adequate joining between the dense parts and porous, reused with permission from Elsevier and John Wiley and Sons [57,60].
Materials 16 03991 g003
Figure 4. MIM stainless steel intravascular stent in (a) as received, (b) folded, and (c) expanded states [59].
Figure 4. MIM stainless steel intravascular stent in (a) as received, (b) folded, and (c) expanded states [59].
Materials 16 03991 g004
Figure 5. MIM processing steps [61].
Figure 5. MIM processing steps [61].
Materials 16 03991 g005
Figure 6. Phase of titanium alloy, reused with permission from Elsevier [81].
Figure 6. Phase of titanium alloy, reused with permission from Elsevier [81].
Materials 16 03991 g006
Figure 7. Yield stress and Young’s modulus for porous titanium compacted toward out-of-plane. Purple zones show the mechanical properties of spongy bone, reused with permission from Elsevier [106].
Figure 7. Yield stress and Young’s modulus for porous titanium compacted toward out-of-plane. Purple zones show the mechanical properties of spongy bone, reused with permission from Elsevier [106].
Materials 16 03991 g007
Figure 8. SEM images of Ti powders: (a) gas atomized, (b) advanced plasma atomized, and (c) HDH, reused with permission from Elsevier [179].
Figure 8. SEM images of Ti powders: (a) gas atomized, (b) advanced plasma atomized, and (c) HDH, reused with permission from Elsevier [179].
Materials 16 03991 g008
Figure 9. MIM samples: (a) unsintered with powder loading 72 vol.%; (bd) sintered at 1200 °C for a period of 3 h with powder loading 72, 75, and 80 vol.%, reused with permission from Elsevier [73].
Figure 9. MIM samples: (a) unsintered with powder loading 72 vol.%; (bd) sintered at 1200 °C for a period of 3 h with powder loading 72, 75, and 80 vol.%, reused with permission from Elsevier [73].
Materials 16 03991 g009
Figure 10. Viscosity versus shear rate at three different temperatures for (a) Feedstock A, (b) Feedstock B, (c) Feedstock C, and (d) Feedstock D, reused with permission from Elsevier [157].
Figure 10. Viscosity versus shear rate at three different temperatures for (a) Feedstock A, (b) Feedstock B, (c) Feedstock C, and (d) Feedstock D, reused with permission from Elsevier [157].
Materials 16 03991 g010
Figure 11. Green parts with powder loading of: (a) 60 vol.% at 165 °C; (b) 53 vol.% at 150 °C; (c) 53 vol.% at 155 °C; (d) 53 vol.% at 160 °C; and (e) 53 vol.% at 165 °C. Reused with permission from Elsevier [68].
Figure 11. Green parts with powder loading of: (a) 60 vol.% at 165 °C; (b) 53 vol.% at 150 °C; (c) 53 vol.% at 155 °C; (d) 53 vol.% at 160 °C; and (e) 53 vol.% at 165 °C. Reused with permission from Elsevier [68].
Materials 16 03991 g011
Figure 12. Injection-molded components in high-pressure applications [19].
Figure 12. Injection-molded components in high-pressure applications [19].
Materials 16 03991 g012
Figure 13. Schematic diagrams demonstrating the distributions of binder in (a) as molded and (b) preliminary stages of solvent debinding on the basis of water extraction, reused with permission from Elsevier [213].
Figure 13. Schematic diagrams demonstrating the distributions of binder in (a) as molded and (b) preliminary stages of solvent debinding on the basis of water extraction, reused with permission from Elsevier [213].
Materials 16 03991 g013
Figure 14. Schematic diagrams demonstrating the distributions of binder in (a) intermediary and (b) final stages of solvent debinding on the basis of water extraction, reused with permission from Elsevier [213].
Figure 14. Schematic diagrams demonstrating the distributions of binder in (a) intermediary and (b) final stages of solvent debinding on the basis of water extraction, reused with permission from Elsevier [213].
Materials 16 03991 g014
Figure 15. Elimination of soluble binder during the process of solvent debinding at different temperatures, reused with permission from Elsevier [68].
Figure 15. Elimination of soluble binder during the process of solvent debinding at different temperatures, reused with permission from Elsevier [68].
Materials 16 03991 g015
Figure 16. Tensile properties of unalloyed titanium versus O content: (a) elongation to fracture, (b) ultimate tensile strength, and (c) yield strength. Samples were produced through powder metallurgy, MIM, and additive manufacturing routes and reused with permission from Elsevier [220].
Figure 16. Tensile properties of unalloyed titanium versus O content: (a) elongation to fracture, (b) ultimate tensile strength, and (c) yield strength. Samples were produced through powder metallurgy, MIM, and additive manufacturing routes and reused with permission from Elsevier [220].
Materials 16 03991 g016
Figure 17. (a,b) SEM image of titanium sintered at 1250 °C. The micron-size pores are indicated by arrows, and (c) energy-dispersive X-ray spectroscopy (EDS) analysis of the scaffold [156].
Figure 17. (a,b) SEM image of titanium sintered at 1250 °C. The micron-size pores are indicated by arrows, and (c) energy-dispersive X-ray spectroscopy (EDS) analysis of the scaffold [156].
Materials 16 03991 g017
Figure 18. (a) Specimen surface after polishing that was sintered at 1250 °C for 2 h; (b) surface microstructure of the similar specimen; (c) at the center of the specimen; and (d) number of pores and average pore size as a function of sintering temperature, reused with permission from Elsevier [46].
Figure 18. (a) Specimen surface after polishing that was sintered at 1250 °C for 2 h; (b) surface microstructure of the similar specimen; (c) at the center of the specimen; and (d) number of pores and average pore size as a function of sintering temperature, reused with permission from Elsevier [46].
Materials 16 03991 g018
Figure 19. Optical micrographs of (a) Ti-8Mn, (b) Ti-9Mn, (c) Ti-12Mn, (d) Ti-13Mn, (e) Ti-15Mn, and (f) Ti-17Mn. Some pores are illustrated by ellipses, reused with permission from Elsevier [165].
Figure 19. Optical micrographs of (a) Ti-8Mn, (b) Ti-9Mn, (c) Ti-12Mn, (d) Ti-13Mn, (e) Ti-15Mn, and (f) Ti-17Mn. Some pores are illustrated by ellipses, reused with permission from Elsevier [165].
Materials 16 03991 g019
Figure 20. SEM fractographs of tensile-tested samples: (a) Ti-8Mn, (b) Ti-9Mn, (c) Ti-12Mn, (d) Ti-13Mn, (e) Ti-15Mn, and (f) Ti-17Mn, reused with permission from Elsevier [165].
Figure 20. SEM fractographs of tensile-tested samples: (a) Ti-8Mn, (b) Ti-9Mn, (c) Ti-12Mn, (d) Ti-13Mn, (e) Ti-15Mn, and (f) Ti-17Mn, reused with permission from Elsevier [165].
Materials 16 03991 g020
Figure 21. Optical micrographs of Ti-16Nb samples sintered at (a) 900 °C, (b) 1100 °C, (c) 1300 °C, and (d) 1500 °C, reused with permission from Elsevier [167].
Figure 21. Optical micrographs of Ti-16Nb samples sintered at (a) 900 °C, (b) 1100 °C, (c) 1300 °C, and (d) 1500 °C, reused with permission from Elsevier [167].
Materials 16 03991 g021
Figure 22. Surface morphology of Ti-6Al-4V/HA composites sintered at (a) 1100 °C, (b) 1200 °C, and (c) 1300 °C, reused with permission from Elsevier [176].
Figure 22. Surface morphology of Ti-6Al-4V/HA composites sintered at (a) 1100 °C, (b) 1200 °C, and (c) 1300 °C, reused with permission from Elsevier [176].
Materials 16 03991 g022
Table 1. Comparison of the mechanical properties of implant materials.
Table 1. Comparison of the mechanical properties of implant materials.
MaterialYoung’s Modulus (GPa)Tensile Strength (MPa)Yield Strength (MPa)Elongation (%)References
Stainless steels and Co-based alloys
316L stainless steel200500–1350200–70010–40[2,18,60,83,109,110,111,112,113,114,115,116,117,118,119]
Co-Cr2005008
Co-Ni-Cr22085020
Type of alloy: α
Pure Ti grade 110324017024
Pure Ti grade 210334527520
Pure Ti grade 310345038018
Pure Ti grade 410455048515
Type of alloy: α + β
Ti-6Al-4V110–114895–9308606–10
Ti-6Al-7Nb114900–1050880–9508.1–15
Ti–5Al-2.5Fe11010207806
Ti-3Al-2.5V10058569015
Type of alloy: β
Ti-13Nb-13Zr79–84973–1037836–90810–16
Ti-35Nb-7Zr-5Ta5559080020
Ti-35Nb-7Zr-5Ta-0.35O41–46876–1015864–9662.2–15.6
Ti-29Nb-13Ta-4.6Zr8091186413.2
Ti-35.3Nb-5.1Ta-7.1Zr5559754719
Ti-24Nb-4Zr-8Sn54–62655–720627–6852–9
Ti-12Mo-6Zr-2Fe74–851060–1100100–106018–22
Ti-15Mo-5Zr-3Al80852–1100838–106018–25
Table 2. Particle sizes of titanium and its alloys used in MIM (years: 2013–2022).
Table 2. Particle sizes of titanium and its alloys used in MIM (years: 2013–2022).
MaterialsParticle Size (µm)References
Ti19.1[73]
<20[74]
74.9[75]
33.2[153]
75[154]
5[155]
<45[156,157,158]
45[159]
26.5[160]
Hydride-dehydride (HDH) Ti45[46]
29[68]
<77[161]
35.43[162]
Ti-12MoHDH Ti: ≤45, hydrogen reduced Mo: ≤25[163]
Ti-Mn<45[164,165]
Ti-NbTi: <45, Nb: <110[166]
Ti: 21, Nb: <75[167]
Ti: 32.95, Nb: 30.15[168]
Ti: <14, Nb: <36[169]
Ti: 21, HDH Nb: 75[170]
Ti-6Al-4V13.4[19]
<45[171]
15[172]
25.5[173]
18[174]
HDH Ti-6Al-4V51.8[47]
70[147]
Ti-6Al-4V/ Hydroxyapatite (HA)Ti-6Al-4V: 19.54, HA: 61.95[175]
Ti-6Al-4V: 19.6, HA: 5[176]
Ti-6Al-4V/Wollastonite (WA)Ti-6Al-4V: 19.54, WA: 10.10[177]
Ti-16Nb-(0-4) SnTi: 23.81, Nb: 14.98, Sn: 20.19[178]
Ti-24Nb-4Zr-8Sn<45[113]
Ti-27.5Nb-8.5Ta-3.5Mo-2.5Zr-5Sn6.08[55]
Table 3. Binders used in MIM processes pertaining to biomedical applications (years: 2013–2022).
Table 3. Binders used in MIM processes pertaining to biomedical applications (years: 2013–2022).
MaterialsBindersReferences
TiParaffin wax, polyethylene, stearic acid[73]
Polyethylene glycol 1500, poly methyl methacrylate, stearic acid[75,154]
Paraffin wax, high-density polyethylene, stearic acid[153]
Polyvinyl butyral, benzyl butyl phthalate, solsperse 20,000[155]
Paraffin wax, high-density polyethylene, stearic acid[156]
Polyethylene glycol, poly methyl methacrylate, polyvinylpyrrolidone[157]
Polyethylene glycol, polypropylene carbonate, poly methyl methacrylate[158]
Polyethylene glycol, poly methyl methacrylate, stearic acid[159]
Polyacetal-based thermoplastic binder[160]
HDH TiParaffin wax, high-density polyethylene, stearic acid[46]
Paraffin wax, low-density polyethylene, stearic acid[68]
Wax-based binders[161]
Agar, sucrose[162]
Ti-12MoLiquid paraffin wax, stearic acid, low-density polyethylene, polypropylene, Polyethylene glycol, naphthalene, solid paraffin[163]
Ti-NbParaffin wax, polyethylene vinyl acetate, stearic acid[167,170]
Paraffin wax, carnauba wax, polypropylene, stearic acid[168]
Paraffin wax, low-density polyethylene, stearic acid[169]
Ti-6Al-4VPolyethylene glycol, polypropylene, stearic acid[19]
Paraffin wax, polyethylene vinyl acetate, stearic acid[171]
Polyoxymethylene, stearic acid, paraffin wax, ethylene vinyl acetate, polyethylene[172]
Paraffin wax, polypropylene, polyethylene, stearic acid[173]
Palm stearin, polyethylene[174]
HDH Ti-6Al-4VPolyethylene glycol, polyvinyl butyral, stearic acid[47,147]
Ti-6Al-4V/HAPalm stearin, low-density polyethylene[175,195]
Palm stearin, polyethylene[176]
Ti-6Al-4V/WAPalm stearin, polyethylene[177]
Ti-16Nb-(0-4) SnParaffin wax, carnauba wax, polypropylene, stearic acid[178]
Ti-24Nb-4Zr-8SnParaffin wax, polyethylene-vinyl acetate co-polymer, stearic acid[113]
Ti-27.5Nb-8.5Ta-3.5Mo-2.5Zr-5SnPolyacetal-based binder[55]
Table 4. Powder loading for titanium and titanium alloys (years: 2013–2022).
Table 4. Powder loading for titanium and titanium alloys (years: 2013–2022).
MaterialsParticle Size (µm)Powder Loading (vol.%)References
Ti19.172, 75, 80[73]
7558[154]
<4569[156]
<4567[157,158]
4560[159]
26.565.6[160]
28.467[173]
HDH Ti4561[46]
2953[68]
<7755[161]
Ti-12MoHDH Ti: ≤45, hydrogen reduced Mo: ≤2565[163]
Ti-NbTi: 32.95, Nb: 30.1550[168]
Ti: <14, Nb: <3660[169]
Ti-6Al-4V13.460[19]
25.564[173]
1865[174]
HDH Ti-6Al-4V51.855, 60[47]
7055, 60[147]
Ti-6Al-4V/HATi-6Al-4V: 19.61, HA: 2068, 69, 70[209]
Ti-6Al-4V: 19.54, HA: 61.9568[175]
Ti-6Al-4V: 19.6, HA: 578.21[176]
Ti-6Al-4V/WATi-6Al-4V: 19.54, WA: 10.1067[177]
Ti-16Nb-(0-4) SnTi: 23.81, Nb: 14.98, Sn: 20.1950[178]
Ti-24Nb-4Zr-8Sn<4565[113]
Ti-27.5Nb-8.5Ta-3.5Mo-2.5Zr-5Sn6.0865[55]
Table 5. Solvent debinding parameters utilized for titanium and titanium alloys (years: 2013–2022).
Table 5. Solvent debinding parameters utilized for titanium and titanium alloys (years: 2013–2022).
MaterialsSolventTemperature (°C)Time (h)References
TiHexane5024[73,153]
Water5035[154]
Hexane5020[156]
Water4028[157]
Water50[159]
HDH TiHexane5020[46]
Heptane45, 60, 756.5[68]
Ti-12MoHeptane5015[163]
Ti-NbHexane4020[166,167]
Heptane608[168]
Heptane5020[169]
Hexane4020[170]
Ti-6Al-4VWater6024[19]
4015[171]
Heptane606[174]
Ti-6Al-4V/HAHeptane606[175]
Ti-6Al-4V/WAHeptane606[177]
Ti-16Nb-(0-4) SnHeptane608[178]
Ti-24Nb-4Zr-8SnHexane4020[113]
Table 6. Thermal debinding parameters for titanium and its alloys (years: 2013–2022).
Table 6. Thermal debinding parameters for titanium and its alloys (years: 2013–2022).
MaterialsThermal Debinding ScheduleReferences
Ti500 °C for 2 h[73,153]
450 °C for 1 h[75,160]
550 °C for 1 h [156]
570 °C for 2 h[158]
HDH Ti550 °C[46]
350 and 550 °C at 3 °C/min[68]
Ti-12Mo700 °C for 2 h[163]
Ti-Nb200–400–600 °C at 2 °C/min for 2 h[168]
250 °C for 2 h and 450 °C for 2 h[169]
Ti-6Al-4V700 °C at 2, 5, and 10 °C/min[173]
550 °C for 4 h[174]
Ti-6Al-4V/ HA500 °C[175]
320 °C at 3 °C/min for 1 h and 500 °C at 5 °C/min for 1 h[176]
Ti-6Al-4V/WA500 °C at 5 °C/min for 1 h[177]
Ti-24Nb-4Zr-8Sn500 °C[113]
Ti-27.5Nb-8.5Ta-3.5Mo-2.5Zr-5Sn500 °C for 2 h[55]
Table 7. Sintering parameters used for titanium and its alloys (years: 2013–2022).
Table 7. Sintering parameters used for titanium and its alloys (years: 2013–2022).
MaterialsSintering Temperature (°C)Time (h)Heating Rate
(°C/min)
Sintering AtmosphereReferences
Ti12003Vacuum[73,153]
12502Argon[75]
13202Argon[154]
125028Argon[155]
1150, 1250, 1300[156]
13002[157]
11502Vacuum[160]
HDH Ti1100–13002Vacuum[46]
135026Argon[68]
11502Vacuum[161]
Ti-Mn11008Vacuum[164,165]
Ti-Nb15004Vacuum[166]
900–150025Vacuum[167]
1500410, 5Vacuum[168]
14004Argon[169]
15002[170]
Ti-6Al-4V13502Vacuum[171]
1200, 1250, 13001, 2, 33, 4, 5Vacuum[174]
Ti-6Al-4V/HA13503Vacuum[175]
1100, 1200, 13002Vacuum[176]
Ti-6Al-4V/WA1100, 1200, 130053Vacuum[177]
Ti-16Nb-(0-4) Sn1250, 1400, 1550210, 5Vacuum[178]
Ti-24Nb-4Zr-8Sn1400, 15002, 4Vacuum[113]
Ti-27.5Nb-8.5Ta-3.5Mo-2.5Zr-5Sn1000, 1100, 1200, 1300, 14008Vacuum[55]
Table 8. Mechanical properties of the sintered Ti and Ti alloy components (years: 2013–2022).
Table 8. Mechanical properties of the sintered Ti and Ti alloy components (years: 2013–2022).
MaterialsSintering Temperature (°C)Young’s Modulus (GPa)Tensile Strength (Mpa)Elongation (%)References
Ti1250, 13007.80, 22[156]
1300617[157]
115099542[160]
HDH Ti125039512.5[46]
Ti-8Mn110087[165]
Ti-9Mn11008910464.7[165]
Ti-12Mn110096[165]
Ti-13Mn110099[165]
Ti-15Mn110098[165]
Ti-17Mn1100103[165]
Ti-16Nb150080667[167]
Ti-Nb1500100[168]
Ti-17Nb140076[169]
Ti-6Al-4V1350824[171]
1200934.33[174]
Ti-6Al-4V + Gd1350749[171]
Ti-6Al-4V/HA130044.26[176]
Ti6Al4V/WA1100–130014.57–18.10[177]
Ti-24Nb-4Zr-8Sn140054656[113]
Ti-27.5Nb-8.5Ta-3.5Mo-2.5Zr-5Sn1100981154[55]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Basir, A.; Muhamad, N.; Sulong, A.B.; Jamadon, N.H.; Foudzi, F.M. Recent Advances in Processing of Titanium and Titanium Alloys through Metal Injection Molding for Biomedical Applications: 2013–2022. Materials 2023, 16, 3991. https://doi.org/10.3390/ma16113991

AMA Style

Basir A, Muhamad N, Sulong AB, Jamadon NH, Foudzi FM. Recent Advances in Processing of Titanium and Titanium Alloys through Metal Injection Molding for Biomedical Applications: 2013–2022. Materials. 2023; 16(11):3991. https://doi.org/10.3390/ma16113991

Chicago/Turabian Style

Basir, Al, Norhamidi Muhamad, Abu Bakar Sulong, Nashrah Hani Jamadon, and Farhana Mohd Foudzi. 2023. "Recent Advances in Processing of Titanium and Titanium Alloys through Metal Injection Molding for Biomedical Applications: 2013–2022" Materials 16, no. 11: 3991. https://doi.org/10.3390/ma16113991

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