Next Article in Journal
Technological Assurance of Surface Roughness of Ti-6Al-4V Parts Made Using Additive and Conventional Manufacturing Methods
Previous Article in Journal
A Bidirectional Digital Twin System for Adaptive Manufacturing
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Processing and Development of Porous Titanium for Biomedical Applications: A Comprehensive Review

by
Mayank Kumar Yadav
1,
Akshay Yarlapati
2,
Yarlapati Naga Aditya
3,
Praveenkumar Kesavan
4,
Vaibhav Pandey
5,
Chandra Shekhar Perugu
6,
Amit Nain
7,
Kaushik Chatterjee
2,
Satyam Suwas
2,
Jayaraj Jayamani
8,9 and
Prashanth Konda Gokuldoss
1,10,11,*
1
Department of Mechanical and Industrial Engineering, Tallinn University of Technology, 19086 Tallinn, Estonia
2
Department of Materials Engineering, Indian Institute of Science, Bengaluru 560012, India
3
Department of Mechanical and Mechatronics Engineering, University of Waterloo, 200 University, Avenue West, Waterloo, ON N2L 3G1, Canada
4
Faculty of Materials Science and Technology, VSB-Technical University of Ostrava, 17. listopadu 2172/15, 70 800 Ostrava, Czech Republic
5
Department of Ceramic Engineering, Indian Institute of Technology (BHU), Varanasi 221005, India
6
Emerging Nanoscience Research Institute (EnRI), Nanyang Technological University, 50 Nanyang Avenue, Singapore 639798, Singapore
7
Department of Applied Mechanics and Biomedical Engineering, Indian Institute of Technology-Madras, Chennai 600036, India
8
Materials Technology, School of Information and Engineering, Dalarna University, 79188 Falun, Sweden
9
Department of Mechanical and Materials Engineering, Karlstad University, 65188 Karlstad, Sweden
10
Centre for Biomaterials, Cellular and Molecular Theranostics (CBCMT), School of Engineering, Vellore Institute of Technology, Vellore 632014, India
11
National Engineering Research Center of Near-Net-Shape Forming for Metallic Materials, South China University of Technology, Guangzhou 510640, China
*
Author to whom correspondence should be addressed.
J. Manuf. Mater. Process. 2025, 9(12), 401; https://doi.org/10.3390/jmmp9120401
Submission received: 31 October 2025 / Revised: 22 November 2025 / Accepted: 1 December 2025 / Published: 4 December 2025

Abstract

Titanium (Ti) and its alloys are widely used in orthopedic applications, including total hip and knee replacements, bone plates, and dental implants, because of their superior biocompatibility, bioactivity, corrosion resistance, and mechanical robustness. These alloys effectively overcome several limitations of conventional metallic implants, such as 316L stainless steel and Co-Cr alloys, particularly with respect to corrosion, fatigue performance, and biological response. However, dense Ti alloys possess a relatively high elastic modulus, which can cause stress shielding in load-bearing applications. This challenge has motivated significant research toward engineered porous Ti structures that exhibit a reduced and bone-matched modulus while preserving adequate mechanical integrity. This review provides a comprehensive examination of powder metallurgy and additive manufacturing approaches used to fabricate porous Ti and Ti-alloy scaffolds, including additive manufacturing and different powder metallurgy techniques. Processing routes are compared in terms of achievable porosity, pore size distribution, microstructural evolution, mechanical properties, and biological outcomes, with emphasis on the relationship between processing parameters, pore architecture, and functional performance. The reported findings indicate that optimized powder-metallurgy techniques can generate interconnected pores in the 100–500 μm range suitable for osseointegration while maintaining compressive strengths of 50–300 MPa, whereas additive manufacturing enables the precise control of hierarchical architectures but requires careful post-processing to remove adhered powder, stabilize microstructures, and ensure corrosion and wear resistance. In addition, this review integrates fundamental aspects of bone biology and bone implant interaction to contextualize the functional requirements of porous Ti scaffolds.

Graphical Abstract

1. Introduction

The demand for orthopedic implants is experiencing an unprecedented surge, primarily due to the increased prevalence of conditions such as osteoporosis, which weakens bones, and osteoarthritis, which causes joint inflammation [1]. Human bone joints are most likely to be affected by degenerative and inflammatory diseases [1]. Degenerative diseases impair the mechanical strength of bones either through excessive stress or due to a lack of the body’s natural healing mechanisms. Studies suggest that nearly 90% of individuals above the age of 40 are affected to some extent by degenerative joint disorders [2]. The replacement of diseased joint surfaces using metallic, ceramic, or polymeric material through arthroplasty surgery is the ultimate solution to this problem [3]. Arthroplasty is a surgical procedure where damaged natural joint surfaces are replaced with specialized implant materials to alleviate pain and improve mobility by creating a new prosthetic joint. Bones comprise four cell populations crucial for growth and resorption: osteoblasts, osteoclasts, osteogenic cells, and osteocytes, all originating from bone marrow. Although bones possess the capacity for physiological remodeling and self-repair, they are unable to effectively cope with severe injuries such as critical-sized defects (CSDs) [4]. CSDs are defects that will not heal on their own during the natural lifespan of the individual [5], and typically, defects in the range of 2 cm to 5 cm are considered critical in humans [6,7]. The standard approaches for treating CSDs include autograft (bone harvested from the same patients), allograft (bone harvested from other patients), and xenograft (bone sourced from other species) [8,9]. While these approaches often yield positive results, they come with specific limitations.
The application of autograft is not suitable for large-sized defects such as spinal arthrodesis [10], while allograft poses risks of immunologic rejection, donor dependency, poor osteogenesis, and a higher resorption rate [11]. Although xenografts present a promising solution to organ shortages, their clinical use is restricted by immune barriers, zoonotic infection risks, ethical concerns, and physiological challenges [12,13,14]. Such challenges and constraints drive researchers to actively explore and create dependable bone substitutes that closely mimic the characteristics of natural human bone. The field of biomaterials research began with the inaugural biomaterial conference held at Clemson University in South Carolina in 1969 [15]. Extensive research has led to the development of various metallic [16], ceramic [17,18], and polymeric [19] materials suitable for mimicking the properties of a healthy human bone. These materials belong to different classes, each possessing distinct physical, mechanical, chemical, and biological properties. For instance, metallic alloys, known for their robust mechanical properties, are commonly used in load-bearing joint prostheses [20,21,22,23]. Ceramic materials, renowned for their wear resistance and bioactive properties, are often used to coat metallic implants [24,25,26,27,28,29]. Polymers, being biodegradable and harmless to the body, are utilized in bone repair applications [30].
Scientists attribute the primary cause of failure in metallic implants to the ‘stress shielding effect’ [31,32,33]. This effect occurs due to the mismatch in mechanical properties like Young’s modulus, yield strength, ultimate tensile strength, fracture toughness, fatigue strength, and ductility (including elongation and toughness under cyclic loading) between natural bone and metallic alloys under in vivo conditions, leading to the resorption of bone tissues surrounding the implant material [34]. Moreover, metallic implants are limited by issues such as poor wear and corrosion resistance, as well as insufficient biocompatibility [35,36,37,38,39,40]. Traditional metallic implants, such as 316L stainless steel and Co-Cr alloys, particularly suffer from localized corrosion, ion release, and inadequate long-term biological compatibility. Although titanium (Ti) and its alloys are also metallic materials, they offer significantly improved corrosion resistance, superior biocompatibility, and better mechanical performance compared to these earlier metallic systems. Owing to their high durability and excellent biological response, Ti-based materials have become the focus of implant development; however, they still present challenges such as a relatively high elastic modulus in dense form, which must be addressed through alloy design and porosity engineering. The commercial production of titanium began with the invention of the Kroll process in 1946 [41], followed shortly by the introduction of titanium alloy–based implants within a year [42]. Since its inception, titanium and its alloys have found extensive use in biomedical applications such as spinal fusion, skeletal repair, and dental implants [15,43,44,45,46,47]. Titanium and its alloys have demonstrated superiority over medical-grade stainless steel, cobalt–chromium, and magnesium-based alloys concerning biocompatibility, mechanical strength, and corrosion resistance. However, their higher Young’s modulus (~110 GPa) initially limited their biomedical applications. The development of second-generation titanium alloys, specifically β alloys, aimed to reduce this modulus. Titanium alloys with a low elastic modulus, such as Ti-Nb-Zr-T, have been specifically developed to closely align with the modulus of different types of human bone [48,49,50,51,52].
The concept of porous biomaterials for osseointegration was introduced by Weber and White in 1972 [53], leading to extensive research on porous implant materials using various processing methods [54,55,56,57]. Porous scaffolds, characterized by a lower effective Young’s modulus than conventional titanium implants, offer several advantages. These materials enhance biological fixation by encouraging bone tissue to grow into the implant’s porous structure, enabling even stress distribution between the implant and surrounding bone. Moreover, porous scaffolds enhance colonization and subsequent substitution by biological cells [58,59]. Numerous manufacturing methods have been developed to produce porous structures for implant applications. Palka et al. [60] reviewed porous titanium implants, emphasizing various types of porosity in their structure, while Koju et al. [61] examined porous Ti6Al4V bone implants produced by additive manufacturing. However, a systematic review encompassing all production methods to fabricate porous implants is still not available. Hence, the primary aim of this review is to survey all existing methods to fabricate porous titanium through powder metallurgy and to provide a comprehensive overview of the research status of titanium and its alloys targeted for orthopedic applications. This article is divided into nine sections with several sub-sections, focusing primarily on the various techniques employed to develop porous titanium scaffolds for orthopedic applications using powder metallurgy.

2. Material Characteristics Critical to Implant Functionality

“As defined by Williams in 1987, biomaterials are non-living materials designed for use in medical devices that interact with biological systems”. Biomaterials are artificial or natural materials that restore the integrity and functionality of damaged or diseased biological structures. For an implant to achieve longevity and high compatibility with the body, the biomaterial must possess certain properties suitable for biomedical applications. These properties include biocompatibility, osseointegration, mechanical strength, wear resistance, and corrosion resistance. Figure 1 outlines the fundamental requirements that must be considered in the design and development of implants.

2.1. Mechanical Properties

The primary function of bone is to provide structural support under mechanical loading and protect vital organs. In cases of trauma or diseased biological structures, implants often replace the function of the affected organ. Currently, most implants used in orthopedic traumas and arthroplasty are of metallic composition due to their excellent mechanical strength [62]; therefore, the mechanical analysis of metallic implants is of prime concern. Essential properties include Young’s modulus, fracture toughness, hardness, yield strength, ultimate tensile strength, stiffness, ductility, time-dependent deformation, and creep. Wolff’s law of bone remodeling (1892) states that mechanical load can affect bone architecture [63] and hence promote bone remodeling and fracture healing [64,65]. Fatigue strength is also critical, as it affects material performance under cyclic loads, influencing the long-term success of implants [15]. Young’s modulus is crucial for the success of implants. A significant mismatch between the modulus of bone and the implant material can negatively impact load transfer from the implant to the bone and within the bone, leading to bone resorption and implant loosening. This biomechanical incompatibility is the stress-shielding effect [15,66,67]. Table 1 lists the mechanical properties of materials used for orthopedic implant applications and human bone.

2.2. Corrosion and Wear Resistance

Corrosion and wear are inevitable problems associated with orthopedic implants [72]. When a metallic implant is introduced into the body, it encounters biological fluid containing various cations and anions, leading to electrochemical corrosion on the implant surface. Four types of corrosion are typically observed in orthopedic implants: galvanic, pitting crevice, and fretting corrosion. Galvanic corrosion occurs due to the electrochemical potential difference between two different or the same metal surfaces when introduced in a biological fluid [73,74,75]. Studies showed that when Ti-based and Co-based alloys are coupled, the corrosion rate was observed to be as low as 0.02 μA/cm2, and no instances of corrosion were found on the metallic interface [76]. Pitting corrosion is a localized form of corrosion that creates cavities in the material [77,78,79]. In the case of metallic implants, chloride ion breakdowns and the protective passive oxide films result in the formation of pits at the site [80,81]. Crevice corrosion is like pitting corrosion in terms of propagation mechanism but differs in the initiation mechanism [82,83,84]. Crevice corrosion generally occurs in confined spaces with low oxygen tension and high chloride concentration, destroying the passivation layer [85]. The mechanism of fretting corrosion differs from other types of corrosion. In fretting corrosion, the passivation layer is disrupted due to micro-motion between parts of an implant [86,87,88]. Continuous relative motion between two surfaces under load generates wear debris around the implant-bone interface. Research indicates that inadequate wear resistance can result in implant loosening, while the accumulation of these wear particles may provoke negative tissue responses [89]. In natural joints, synovial fluid functions as an effective lubricant, providing boundary and mixed lubrication regimes that significantly reduce wear [90]. However, the stability of this lubricant film depends strongly on surface chemistry, roughness, and pore geometry, which regulate protein adsorption and fluid retention. Poorly designed surface features or trapped debris can disrupt the synovial film and accelerate third-body abrasion. Therefore, implant materials and surface designs must ensure adequate corrosion and wear resistance while also supporting stable synovial lubrication to prevent premature mechanical failure.
Corrosion in metallic implants is an electrochemical degradation process that occurs when the metal interacts with physiological fluids. As described in the review article by Eliaz et al. [91], the human body behaves as an electrolyte containing chloride ions, proteins, and dissolved oxygen, which facilitates metal dissolution through anodic and cathodic reactions. The breakdown of the passive oxide film, especially under mechanical loading or fretting conditions, results in the release of metallic ions and particulate debris into surrounding tissues. This degradation can lead to local inflammatory responses, hypersensitivity, cytotoxic effects, and ultimately implant loosening or failure. The article emphasizes that corrosion is strongly influenced by alloy composition, microstructure, mechanical wear, and the stability of the protective oxide layer on the implant surface. Therefore, selecting materials with high corrosion resistance, such as titanium and its alloys, and optimizing surface treatments are essential for minimizing clinical complications associated with corrosion-induced degradation in biomedical implants [91].
The corrosion mechanism inside the human body can be easily understood by taking the example of Fe and Ti in biological environments. In physiological (aqueous) environments, the corrosion of metallic implants is propelled by coupled electrochemical half-reactions: metal dissolution (anodic) and reduction (cathodic) at the metal/electrolyte interface. For titanium alloys (e.g., Ti-6Al-4V), corrosion begins with the oxidation of titanium atoms, producing Ti4+ ions that hydrolyze to form a stable titanium dioxide (TiO2) passive film. This passivation layer strongly inhibits further metal dissolution under steady-state conditions, but local disruption (due to mechanical stress, chloride, or reactive oxygen species) can lead to breakdown and renewed dissolution. Studies show that inflammation (e.g., H2O2) and cathodic activation can degrade the oxide film and drive selective dissolution of phases, particularly the β-phase in Ti-6Al-4V. The relevant electrochemical reactions can be expressed as follows:
For titanium alloys, we have the following:
Anodic Reactions (Oxidation)
T i T i 4 + + 4 e
Passive Film Formation
T i 4 + + 2 H 2 O T i O 2 + 4 H
This is consistent with the mechanism of oxide-film stabilization seen in studies of Ti-6Al-4V [92].
Cathodic Reaction (Reduction)
O 2 + 2 H 2 O + 4 e 4 O H
This is the dominant cathodic process in neutral, oxygenated physiological solutions [93].
For Fe-based alloys, corrosion typically proceeds via continuous dissolution due to the relatively less stable protective films in chloride-rich physiological environments.
The primary anodic reactions are as follows:
F e F e 2 + + 2 e
F e 2 + F e 3 + + e
Meanwhile, the cathodic reaction is generally the same oxygen-reduction process:
O 2 + 2 H 2 O + 4 e 4 O H
These equations and their relevance to in vivo degradation have been documented in the context of biodegradable metal implants [93,94].

2.3. Porosity Effect

Porosity is crucial in developing effective implants for bone tissue engineering. The key factors include pore interconnectivity, porosity volume, and pore sizes. The porous structure must have interconnected pores of appropriate size [95,96]. Interconnected pores are vital as they facilitate cell infiltration, in vivo vascular formation, osteogenesis, cell proliferation, oxygen and nutrient flow, and waste removal [97,98]. It is also reviewed that interconnectivity provides a way for the ingrowth of blood vessels to facilitate multiple vessel formation [99]. Earlier studies suggest that a minimum interconnection size of over 50 µm is suitable for bone ingrowth [100]. The porosity of the implant enhances bone integration because the porous surface provides an interlocking medium to the surrounding tissue and implant, resulting in good biomechanical compatibility and high resistance to fatigue loading [101,102]. It has been reported that pore sizes ranging between 100 µm and 200 µm are suitable for osseointegration [15].
It is important to note that any increase in the implant’s porosity content decreases the mechanical properties of the system [103]. Therefore, without compromising the mechanical properties, the porosity of the system should be optimized, and the pore size should be strictly controlled. Different techniques for adding space holder materials of defined size, densification of the green compacts, sintering conditions, etc., can be employed to incorporate pores in the final structure [104]. A detailed discussion of the production of porous structures using different techniques will be discussed in subsequent sections.

2.4. Surface Wettability and Its Role in Implant Performance

Surface wettability is a critical design parameter for metallic implants, directly influencing both their biological and tribological performance. Wettability governs the interaction of synovial fluid, proteins, and cells with the implant surface, thereby affecting lubrication, friction, wear, and early-stage osseointegration. Hydrophilic surfaces generally facilitate more favorable biological interactions compared to hydrophobic surfaces, largely because wettability affects the first biomolecular events at the implant interface. Surface wettability is known to regulate four major aspects of host response: (i) adsorption and conformation of proteins and macromolecules that form the initial conditioning layer; (ii) adhesion, spreading, and differentiation of hard and soft tissue cells; (iii) bacterial adhesion and biofilm formation; and (iv) the overall rate and quality of in vivo osseointegration [105]. These interactions begin within milliseconds of implantation and significantly depend on the hydrophilic or hydrophobic nature of Ti surfaces.
Recent works have demonstrated that micro- and nano-scale surface textures strongly influence the spreading characteristics and retention of lubricants and biological fluids on Ti alloys. Several studies [106,107] highlight that multiscale fractal roughness enhances dynamic wetting, fluid transport, and boundary lubrication, thereby reducing wear at articulating implant interfaces. These findings emphasize that wettability is not only important for osseointegration but also for the flow behavior of synovial fluid, which plays a major role in limiting wear during joint articulation. Insights from Gittens et al. [108,109,110] further confirm that hydrophilic surfaces support improved protein adsorption in favorable conformations, enhance osteoblast maturation, and promote faster bone–implant contact compared to hydrophobic surfaces. Hydrophilic and superhydrophilic surfaces also reduce hydrocarbon contamination and strengthen the biochemical environment for cell attachment, proliferation, and differentiation [111]. Conversely, hydrophobic surfaces may denature adsorbed proteins, inhibit cell adhesion, and alter the biological cascade leading to integration. Overall, incorporating appropriate surface wettability, often achieved through micro/nano-texturing, plasma modification, chemical treatments, or photocatalytic activation, provides a significant advantage for titanium implants [112]. It enhances lubrication and wear performance in load-bearing joints and concurrently improves osseointegration, offering a dual benefit highly relevant to porous titanium structures discussed in this review.

2.5. Biocompatibility

The biocompatibility of any material is the basis of understanding the host response to implants and is defined diversely by different researchers and regulatory bodies. According to the US Food and Drug Administration, biocompatibility is defined as the effect that the materials induce no measurable harm to the host [113]. It is observed that improper biocompatibility leads to dysesthesia (loss of sense), discomfort, pain, infection, resorption of bone [114], etc. The ions released from the metallic implant may induce hypersensitivity and lead to implant failure [115]. Williams (1987) [116] defines biocompatibility as a material’s capacity to generate a suitable response from the host in a given medical application. This implies that a biocompatible material must exhibit both bioactivity and bifunctionality.
The biocompatibility of any material can be defined as the ‘ability of a material to perform with an appropriate host response in a specific application’, which means that a biocompatible material should be bioactive and biofunctional. Similarly, Retner et al. [117] defined that no material can be biocompatible if it leaches cytotoxic substances when implanted. The success of biomaterials largely depends on the body’s reaction to the implant, which measures the material’s biocompatibility. In other words, an implant material is considered biocompatible if it elicits a positive response when exposed to a biological environment [118]. Biocompatibility is primarily determined by two key factors: the biological response the material triggers in the host and the extent to which the material degrades within the bodily environment [15]. Therefore, to achieve good biocompatibility, non-toxic alloying elements with excellent corrosion and wear resistance should be chosen during the design and development of metallic implants.
As discussed in Section 2.2, corrosion can directly affect the overall success rate of the implant, as biocompatibility and corrosion are intrinsically linked. As the degradation of metallic implants in vivo directly influences the host response. A biocompatible material must not trigger thrombosis, toxic or allergic reactions, adverse immune responses, or destruction of cellular components, and it should maintain its structural integrity without releasing harmful ions into surrounding tissues [91]. However, corrosion alters the implant surface chemistry and accelerates metal-ion release, which can destabilize local biological equilibrium and initiate inflammation or hypersensitivity. As corrosion progresses, ions such as Ni, Cr, Co, or V can accumulate and provoke cytotoxic or allergenic responses, sometimes even at low corrosion rates that do not compromise mechanical performance [91].
Body fluid composition further governs this relationship. Chloride ions promote passive-film breakdown, fluctuating pH during inflammation (as low as 4.0) increases localized corrosion, and reduced dissolved oxygen levels hinder repassivation, collectively intensifying degradation and influencing biocompatibility outcomes [119]. Proteins and macromolecules in extracellular fluids can bind metal ions and transport them away from the surface, destabilizing the electrical double layer, or alternatively form an adsorbed barrier that slows corrosion; thus, protein–metal interactions directly affect degradation rates and biological responses. Cells adhering to implant surfaces may either protect the metal by forming a physical barrier or enhance corrosion through secretion of reactive oxygen species such as O2− and H2O29, which accelerate metal dissolution and negatively impact surrounding tissue [91].

2.6. Osseointegration

According to the American Academy of Implant Dentistry (1986), osseointegration refers to the direct contact between normal remodeled bone and an implant without any intervening non-bone tissue, ensuring sustained load transfer and distribution from the implant to and within the bone tissue. Clinically, osseointegration can be defined as the process by which a clinically asymptomatic, rigid fixation of alloplastic materials is achieved and maintained in bone under functional loading [120]. Factors such as design, surface chemistry, roughness, chemical composition, and loading conditions should be considered before developing an implant for optimal osseointegration [121]. Figure 2 shows a schematic diagram illustrating the timeline of osseointegration of a metallic implant with respect to cellular events for 28 days [122].

3. Bone Metabolism

Bone is a rigid, living tissue that provides support during mechanical loading and protects vital organs inside the body [123]. When a baby is born, their body contains approximately 300 soft bones. These bones are primarily made of cartilage, which allows for flexibility and growth. As the baby matures, bones gradually fuse and harden through ossification. By adulthood, this fusion reduces the total number of bones to 206, providing a more rigid and supportive skeletal structure. To understand bone metabolism, it is necessary to grasp bone physiology and mechanical properties. The different structural levels of bone are shown in Figure 3, illustrating that the bone tissue comprises collagen fibers at the nanoscale.

3.1. Bone Physiology

At the microscopic level, bone structure can be classified into two types based on the arrangement of fibers: woven bone and lamellar bone [124]. The following sections will explore these two types in detail.

3.1.1. Woven Bone

Woven bone, considered an immature type of bone tissue, typically has a mineral grain size ranging from 10 to 15 nanometers. It is commonly located in the metaphyseal region and at sites of fracture healing, such as the callus [125]. These are coarse-fibred and collagen fibers and are randomly oriented throughout the structure. It can also be inferred that the direction-independent mechanical behavior of woven bone results from nonuniform collagen [126].
Figure 3. (a) Macro-to-microscopic picture of cancellous and cortical bones. The endosteum structure lines the spaces of cancellous bone, where bone marrow is stored. The periosteum barrier protects the cortical tissue that is tightly packed within osteons. Haversian canals, which comprise blood vessels and nerve tissue, surround concentric lamellae with thicknesses of around 3 μm to create osteons. Osteocytes live in the osteon’s lacuna structures. (b) At the nanometric scale, bone tissue comprises collagen fibers with 67 nm periodic spacing and 40 nm gaps for mineral components. Reproduced with permissions from [127].
Figure 3. (a) Macro-to-microscopic picture of cancellous and cortical bones. The endosteum structure lines the spaces of cancellous bone, where bone marrow is stored. The periosteum barrier protects the cortical tissue that is tightly packed within osteons. Haversian canals, which comprise blood vessels and nerve tissue, surround concentric lamellae with thicknesses of around 3 μm to create osteons. Osteocytes live in the osteon’s lacuna structures. (b) At the nanometric scale, bone tissue comprises collagen fibers with 67 nm periodic spacing and 40 nm gaps for mineral components. Reproduced with permissions from [127].
Jmmp 09 00401 g003

3.1.2. Lamellar Bone

Lamellar bone is another structure of bone that can be distinguished at the microscopic level. Unlike woven bone, it consists of a mineralized matrix known as hydroxyapatite, with the chemical formula Ca10(PO4)6(OH)2. In the human femur, hydroxyapatite crystals typically measure between 20 and 80 nm in length and from 2 to 5 nm in thickness [125]. As implied by their name, lamellar bones are structured in layers and feature collagen fibers aligned with stress directions, contributing to their anisotropic mechanical behavior [126].
Further, based on structural organization, woven and lamellar bones are organized into trabecular and cortical bones [125,128]. Trabecular bone is highly porous, having 50 to 90% porosity and large pores up to several millimeters in diameter. Therefore, these bones are also known as spongy bones or cancellous bones. The trabecular bone bears compressive forces under physiological loading conditions [126]. These bones are typically located in the metaphyseal and epiphyseal regions of both long and cuboidal bones. The distal end radius is an example of a trabecular bone. Similarly, cortical bones are less porous and possess small pores of size up to 1 mm in diameter. Therefore, it is also known as compact bone, contributing to 80% of the weight of the human skeleton. It is harder, stronger, and stiffer than trabecular bone due to less porosity. The humerus and femur are examples of cortical bone.

3.2. Chemical Composition of Bone

Bone is primarily composed of approximately 70% inorganic material, 20% organic content, and 10% water. The inorganic portion mainly consists of crystalline hydroxyapatite (HAp) Ca10(PO4)6(OH)2, while around 90% of the organic component is made up of Type I collagen. The remaining 10% includes non-collagenous proteins, lipids, and various other macromolecules [125,126]. HAp is the principal inorganic component of natural bone and has been widely studied for orthopedic and dental applications because of its exceptional biocompatibility and bioactivity. Its crystal structure is hexagonal, and its Ca/P molar ratio of 1.67 closely matches that of biological apatite, enabling strong chemical affinity with native bone tissue [17,18]. Synthetic HAp exhibits the ability to directly bond with bone through the formation of a biologically active apatite layer, which stimulates osteoblast adhesion, proliferation, and mineralization. In addition to its osteoconductive behavior, HAp provides good compressive strength and structural stability, although it remains brittle and exhibits relatively low fracture toughness, which limits its role as a standalone load-bearing material. Consequently, HAp is frequently used as a coating material to enhance osseointegration and improve the biological interface. Its thermal stability, ionic substitution capability (e.g., carbonate, magnesium, zinc), and similarity to natural bone mineral make HAp an essential material for engineered bone scaffolds and composite implant systems. Boskey and Coleman [129] suggested that the mineral component of bone changes with age. Various changes in bone mineral composition, which occur with age, are listed below [123]:
  • An increase in overall mineral content;
  • Greater carbonate substitution within the mineral structure;
  • A reduction in acid phosphate substitution;
  • Higher hydroxyl content;
  • An elevated calcium-to-phosphorus (Ca/P) molar ratio;
  • Growth in crystal size and improved crystallinity.

3.3. Types of Cells in the Bone

Bones comprise four types of cells: osteoblasts, osteoclasts, osteocytes, and osteogenic cells. Figure 4 provides a schematic diagram of these cell types [130].

3.3.1. Osteoblasts

These cells are also known as bone-forming cells because they synthesize new bones. Osteoid (a protein mixture secreted by osteoblasts) plays an important role in this process, as it mineralizes to form bone [130].

3.3.2. Osteocytes

Osteocytes are fully differentiated bone cells derived from osteoblasts. The primary distinction between these two cell types lies in their position within the bone structure: osteoblasts reside along the periosteal and endosteal surfaces, while osteocytes are embedded within the bone matrix, organized in concentric rings around the central canal of an osteon and situated between lamellae [130].

3.3.3. Osteoclast

Osteoclast cells are responsible for bone tissue breakdown and essential for maintaining, remodeling, and repairing the vertebral skeleton’s bones. The osteoclast secretes acid and collagenase through bone resorption, which molecularly breaks down and digests the mixture of hydrated proteins and minerals. This mechanism similarly regulates the blood calcium level. A bone with a physiological grain size of less than 100 nm in diameter is generally termed a healthy bone. A healthy bone is continuously remodeled throughout the life span by a process that involves the formation of a bone modeling unit and activating bone cells. Osteoclasts perform the phenomenon of bone resorption. The osteoclast is derived from pluripotent cells of the bone marrow. Pluripotent cells tend to discriminate against different cells, including monocytes and macrophages. These pluripotent cells resorb bone by forming disordered cell membrane edges and thus increase the surface area of attachment on the bone surface. The carbonic anhydrase system produces hydrogen ions, which decreases the pH of the local environment, and due to this, the solubility of hydroxyapatite (a major inorganic component of bone) increases. After this process, proteolytic digestion takes place to remove an organic component, and the removal of organic and inorganic components results in the formation of resorption pits [131].

3.3.4. Osteogenic Cells

Osteoprogenitor cells, also known as osteogenic cells, are stem cells found in bone tissue that play a vital role in bone formation and regeneration. Located within the bone marrow, they serve as precursors to more specialized bone cells, including osteoblasts and osteocytes. From immature mesenchymal cells, osteoprogenitor cells develop into spindle cells on the surface of adult bones. They are more common in growing bones, triggering multifunctional phases that reshape them. As we age, our body’s capacity to produce or use more osteoprogenitor cells declines [132].

4. Materials Used for Orthopedic Implant Applications: Advantages and Disadvantages

Materials such as metals, polymers, and ceramics are commonly used for implants. This review article focuses on an in-depth study of the development of Ti-based porous materials. Still, it also includes a brief section on other materials like ceramics and polymers used in implant applications. The selection of material depends on the application area. Table 2 briefly describes the advantages, disadvantages, and applications of various materials used as implants. Ceramics are inorganic oxides of non-metals known for their excellent corrosion resistance, biocompatibility, and bioactivity. They are primarily composed of ionic bonds with some covalent character, making them brittle with low fracture toughness and a high elastic modulus. There are two main types of bioceramics: bio-inert ceramics and bioresorbable ceramics. Bio-inert ceramics, such as zirconia and alumina, exhibit high chemical inertness, making them suitable for biological applications [133,134,135]. Bioresorbable ceramics, like hydroxyapatite and β-tricalcium phosphate, are designed to gradually dissolve and be replaced by natural bone [136,137,138]. However, due to their poor mechanical properties, pure ceramics are not ideal for weight-bearing prostheses in long bones, as they tend to fail under load-bearing conditions. However, ceramics are successfully used in bone filling and dentistry applications [139]. Similarly, polymers are organic non-metal oxides; they can be used as an implant material in low-load-bearing fracture sites. Polymers have emerged as promising materials for orthopedic use due to their mechanical properties, which are comparable to those of trabecular bone, and their biodegradability has further heightened interest in their application [140]. Polymer-like ultra-high-molecular-weight polyethylene (UHMWPE) has properties like high impact strength, low friction coefficient, and low density. These properties have made UHMWPE a popular choice for joint replacement. However, the application of UHMWPE is limited due to long-term radicals in the bulk resulting from the ionizing radiation employed in the sterilization process [141]. Natural polysaccharide polymers like starch, cellulose, and alginate are also used in biomedical applications [139,140]. The main problem associated with the use of polymers in orthopedic applications is the overproduction of wear debris, which leads to inflammatory reactions between adjacent tissue and implant. This adverse tissue reaction causes osteolysis, bone resorption, and implant failure [142].
Metals and their alloys are preferred over ceramics and polymers for implant applications due to their excellent mechanical properties and biocompatibility [143]. Commonly used metals include titanium and its alloys, medical-grade stainless steel, cobalt–chromium, and magnesium alloys. These metals may contain elements such as Co, Cr, Al, Cu, V, and Ni, classified as allergic. Additionally, elements like Cd, Be, Pb, Ba, and Th are classified as toxic [144,145]. These considerations are crucial in ensuring the safety and efficacy of implants in medical applications. Elements like Ni, Co, and Cr are released from 316L stainless steel and Co-Cr alloy due to corrosion in the body environment [146,147]. The toxicity of Ni causes skin-related diseases like dermatitis, and the release of Co causes carcinogenicity [15]. Young’s modulus of 316L SS and Co-Cr alloy is much higher than that of natural bone, causing nonuniform load transfer between bone and implant and leading to bone resorption and implant loosening. Also, 316L SS steel has poor fatigue strength and wear resistance, limiting its application in orthopedics. 316L exhibits acceptable bulk corrosion resistance but is susceptible to localized pitting and crevice corrosion in chloride-rich physiological fluids and may release Ni/Cr ions under certain conditions, leading to allergic responses.
However, due to the low cost of 316L SS compared to all other metallic alloys, 316L SS has maintained its demand in fixation devices like bone plates, bone screws, etc. Magnesium, zinc, and their alloys, due to their biodegradability and potential for preventing revision surgery, have increased the attention of the general orthopedic community for surgical fixation of injured musculoskeletal tissue [148,149,150,151]. It is believed that when a Mg-based alloy is introduced in a saline environment, it degrades to magnesium chloride, oxide, sulfate, or phosphate, and these ions do not cause any adverse effect on local tissues [152,153]. Despite many advantages, there are certain limitations to using Mg-based alloys. The high corrosion rate of Mg-based alloys is one of the major problems [154]. Corrosion causes the evolution of hydrogen gas, which creates the balloon effect in vivo [155]. Due to the high corrosion rate, the pH value of the surrounding surface also increases [156].
Table 2. Table shows a comprehensive list of biomaterials, along with their advantages, disadvantages, and applications [15,17,157].
Table 2. Table shows a comprehensive list of biomaterials, along with their advantages, disadvantages, and applications [15,17,157].
MaterialAdvantagesDisadvantagesApplications
SS 316LWidely available and cost-effective, excellent mechanical properties, biocompatibleHigh elastic modulus, inadequate resistance to corrosion, low wear resistance, potential to trigger allergic reactions in surrounding tissues, and stress shielding, which can lead to bone resorptionBone plates, bone screws, pins, wires, etc.
Co-Cr alloysExcellent resistance to corrosion, fatigue, and wear. High mechanical strength. Sustained biocompatibility over the long termHigh cost, limited machinability, induction of stress shielding, potential biological toxicity from the release of cobalt (Co), chromium (Cr), and nickel (Ni) ionsShorter-term implants, bone plates and wires, total hip replacements (THR), and stem or hard-on-hard bearing system
Mg alloyBiocompatible, biodegradable, bioresorbable, similar density, Young’s modulus is that of natural bone, less stress-shielding effect, and lightweightHydrogen evolution during degradation and less corrosion resistanceBone screws, bone plates, bone pins, etc.
Ti alloyExcellent resistance, lower modulus, stronger than stainless steel, lightweight, and biocompatiblePoor wear resistance, poor bending ductility, and expensiveFracture fixation devices such as plates, nails, rods, screws, fasteners, and wires; femoral hip stems; total koint replacement (TJR) systems; and arthroplasty procedures, particularly for hip and knee joints
Alumina (Al2O3)Biocompatibility and bio-inert behavior, elevated hardness, strength, resistance to abrasion, minimal formation of fibrous tissue at the implant–tissue interfaceLow fracture toughness, brittleness, limited ductility, and radiopacityPorous coatings for femoral stems, femoral head, bone screws and plates, and knee prosthesis
Zirconia
(Zr2O3)
Excellent fracture toughness; high flexural strength; low Young’s modulus; closely matching that of bone; bio-inert nature; good biocompatibility; and non-toxic behavior within the biological environmentPhase transformation, brittleness, and low toughnessFemoral head, artificial knee, bone screws, and plates
BioglassBiocompatibility, bioactivity, promoting integration with surrounding tissue, non-toxicity, and brittleness, which may limit load-bearing applicationsBrittleness, low tensile strength, and poor fatigue resistanceArtificial bone and dental implants
Hydroxyapatite (HAp)Bio-resorbable, bioactive, biocompatible, similar composition to bone, and good osteoconductive propertiesBrittleness, low tensile strength, and poor fatigue resistanceFemoral knee, femoral hips, tibial components, and acetabular cup

5. Titanium and Its Alloys: Material of Ultimate Choice for Implant Application

Titanium (Ti) is the ninth most abundant element in the lithosphere, as it is a constituent of practically all crystalline rock. Reverend William Gregor discovered it in 1798 [114]. Thorough explanations of titanium’s science, technology, and applications may be found in recent investigations. During the last 30 years, titanium and its alloy manufacturing techniques have advanced at a faster rate than any other structural material in metallurgical history. To enhance the performance and dependability of Ti and its alloys and prevent unanticipated failures, there has been an exceptionally high focus on removing manufacturing errors, from extraction and melting procedures to machining and joining. We observe that the primary impediment to an even wider variety of Ti and its alloy applications remains costly. The need to convert Ti oxides or chlorides to their metallic state accounts for a significant portion of the material’s cost. The industrial reduction technique is a batch operation that requires a lot of energy. It was developed many years ago by Kroll. Several initiatives are underway to co-reduce mixed chlorides using various reduction techniques to create metallic or alloyed Ti. Several of these techniques have proven workable in the lab, but scaling up to manufacturing volumes is proving more difficult than anticipated.
Ti is a transition metal with atomic number 22 with an incomplete valence shell. It can form a substitutional solid solution (SSS) with an element having a size factor of ±20%. The melting point of Ti is ~1678 °C, and the crystal structure of Ti is a hexagonally closed-packed (hcp) α-Ti structure up to the beta transus temperature (882.5 °C), transforming to a body-centered cubic β-Ti structure (bcc) above this temperature (Figure 5) [158,159]. The nature of the alloying element determines the microstructure (alpha (α) to beta (β) transformation temperature) of the Ti-based alloy. Elements like Al, O, N, etc., are commonly known as α-stabilizers because they tend to stabilize the α phase, as adding these elements increases the β transus temperature. Similarly, elements like V, Mo, Nb, Fe, Cr, etc., stabilize the β transus temperature and are known as β stabilizers. The addition of this element depresses the β transus temperature.
Titanium alloys are generally classified based on alpha (α) and beta (β) phases. They are classified as α, near-α, (α + β), and metastable β-alloy. The α alloy consists of only α phase (with or without the help of α stabilizers, i.e., Al, O, N, etc.). Near α alloys are a special class of α alloys comprising 1–2% of β stabilizers and about 5–10% of the β phase. Similarly, α + β alloys contain 10–30% of the β phase, and alloys with higher β stabilizers where the β phase is formed by fast cooling are known as metastable β alloys. Generally, α + β or metastable β alloys are employed in biomedical applications. Among all these alloys, α and (α + β) alloys are considered first-generation Ti alloys and possess a high Young’s modulus value (110 GPa). The development of a second-generation Ti-based alloy, i.e., β alloy, was established in 1990 [160]. Due to their ability to possess a lower Young’s modulus value (55–90 GPa), β alloys are the material of choice for orthopedic applications. The different Ti alloys and the mechanical properties used for biomedical applications are mentioned in Table 3. Figure 6 highlights the various biomedical applications where Ti and its alloys are used.

6. Fabrication of Porous Titanium Using Various Powder Metallurgical Techniques

Ti-based materials have low thermal conductivity and high reactivity with the surrounding environment. Due to this, their machining, melting, and casting are difficult. Therefore, Ti-based components are generally machined from forged Ti blanks at a low speed; in this procedure, ~95% of the raw materials are lost as scrap, and recycling this scrap is still a challenge [161]. Also, dense Ti-based alloys tend to have a stress-shielding effect due to their high modulus. Incorporating pores is a promising solution to reduce the stress-shielding effect in Ti-based implants, but manufacturing porous Ti-based structures is not technically easy. Various manufacturing methods have been established to produce porous titanium structures. Figure 7 presents a graphical comparison of porosity levels and pore size distributions achieved through these different fabrication techniques [160]. The use of porous material in artificial joint replacement is an attractive field of research as it includes different methods and materials that can be used to reduce stiffness mismatch with bone. In the present study, different methods for the synthesis of porous Ti scaffolds are described. Figure 8 shows the classification of various techniques that can be used to fabricate porous Ti structures. However, the following sections will focus exclusively on a few processes that can fabricate porous Ti-based structures.

6.1. Space Holder Technique

The space holder technique is one of the important powder metallurgical techniques that can control the size and shape of the pores, amount of porosity, etc. Because these parameters depend on the size of space-holder particles. The basic requirement of this process is that the particle size of the metal powder should be less than the particle size of the space holder. This method involves the addition of space holder particles with the metal powder, followed by mixing them uniformly throughout. Then, the mixture is compacted uniaxially to form a green compact. The green sample is pre-sintered at an optimized temperature (low temperature) so that the complete removal of space holder particles can occur. This also leads to the initial sintering of metal powders. Finally, the pre-sintered samples are sintered at an elevated temperature in an inert atmosphere to avoid contamination (including metal oxidation). The schematic representation of the process for preparing porous Ti via the space holder technique is shown in Figure 9. Several materials have been used as space holders, including bio-wastes, metals like Mg granules, urea, ammonium hydro carbonate, water-soluble materials (like sucrose, potassium chloride, and sodium chloride), paraformaldehyde, etc. Before selecting the space holder, the factors to be considered are its affinity with Ti, the amount of left-out residue after burning, and ease of processing. Complete removal of space holder material from the substrate is the main problem associated with the space holder technique because the presence of any residue may impart detrimental effects, which may influence biocompatibility. Materials like NaCl and sucrose are suggested as they can be removed completely when treated with water [162].
Kim et al. [163] used sacrificial Mg granules as space holder particles since Mg ions are directly involved in numerous biological mechanisms in our body, such as the channelizing of ions, DNA stabilization, enzyme activation, and stimulation of cell growth and proliferation. So, the problems associated with space holders, like partial removal of space holder particles, are completely mitigated. Mg particles of size 20 mesh to 100 mesh with 0.5 wt.% ethanol as a binder were used. The mixture of Ti and Mg was compacted uniaxially, followed by removing space holder particles by dipping them in HCl and ethanol for 24 h. Sintering the green samples at an elevated temperature of 1300 °C for 2 h in high vacuum results in controlling the porosity within 50–71% and pore size within 132–262 µm. The mechanical testing of the samples reveals that the compressive and rupture strengths are 59–280 MPa and 85 MPa, respectively. Esen and Bor [164] processed Ti foam using Mg as a space holder particle; the porosity content was observed to vary between 45 and 70%; the pore size was found to be 525 µm; Young’s modulus ranged between 0.42 GPa and 8.8 GPa.
Bio-waste, like rice husk, is a rich source of silica, and its low-temperature combustion property can be utilized for space-holder applications. Due to the amorphous nature of silica and its high relative reactivity along with carbon for thermal reduction, rice husk can be a favorable candidate for synthesizing low-temperature porous material [165,166]. Apart from the above space holders, researchers have used many other materials to produce porous titanium scaffolds. Dabrowski et al. [3] employed paraformaldehyde, with an average particle size of 500 µm, as a space holder in the fabrication of porous titanium implants due to its ability to fully decompose at relatively low temperatures. The resulting porosity can vary between 60% and 70%, and Young’s modulus can vary between 1 GPa and 8 GPa, which is very close to that of cancellous bone. Xiao Jian et al. [167] studied the porous Ti structures produced with varied spacer contents (Figure 10).
Reports suggest that urea is an effective space-holder material that tends to decompose at low temperatures, resulting in the formation of a pore. Vasconcellos et al. [168] synthesized porous Ti with 3D interconnected pores with a size of ~480 µm and a total porosity of 36%. Wenjuan et al. [169] used urea of size 200–600 µm as a space holder and polyethylene glycol as a binder to synthesize a porous Ti scaffold with porosity varying between 55 and 75% and pore size varying between 200 µm and 500 µm. The porous scaffold reveals Young’s modulus in the range of 3–6.4 GPa, which is close to that of natural bone. It has also been noticed that the rapid decomposition of urea at low temperatures causes rough control over porosity. The needle-like shape of urea particles provides sharp corners and notches in the pore, which lead to stress concentration, deteriorating their mechanical properties. It is also found that the remaining urea in pores makes the implant unfit for medical use [170].
Xiang et al. [171] prepared porous Ti with a porosity between 44 and 77% by using ammonium acid carbonate as a space holder. The fabricated porous Ti scaffolds have pore sizes between 200 µm and 500 µm, with Young’s modulus and compressive strength values ranging between 2.1 GPa and 3.4 GPa and 60 MPa and 140 MPa, respectively. In addition to the space holders, other materials can serve as pore formers, meeting essential criteria such as complete residue removal through dissolution in water and being economically viable. In this class, sucrose and NaCl are the best-suited materials that can be used as a pore former. Torres et al. [172] reported excellent adhesion and proliferation of cells when NaCl is used as a space holder. Also, any left-out residue in the scaffold will not affect the in vivo performance of the Ti alloy.
Chen et al. [173] introduced a novel space holder material that offers excellent biocompatibility for fabricating porous titanium structures with open and interconnected pore networks. Spherical sugar pellets were used to synthesize porous Ti with 20 to 54% porosity and a pore size ranging from 212 µm to 500 µm. The Young’s modulus of the scaffold is found between 12.1 GPa and 18.5 GPa, which is very close to that of natural bone. Polymethyl methacrylate (PMMA) is also used as a space holder material to fabricate porous scaffolds. Li et al. [174] used PMMA to produce macro pores of size 200–400 µm and porosity 10–65%. The green compact was heated between 250 and 450 °C to completely remove the space holder particles. The compressive strength and elastic modulus were observed to be between 32 MPa and 530 MPa and 0.7 GPa and 23.3 GPa, respectively. Table 4 summarizes the mechanical properties of different porous materials.

6.2. Replication Method

The synthesis of porous Ti with the help of replicating polymeric sponges followed by high-temperature sintering is a unique technique. This process offers the fabrication of scaffolds with a high degree of porosity and highly interconnected microspores with identical shapes and sizes. The porous structure produced via this method has a pore shape and size like cancellous bone [178,179,180,181]. In this method, the precise regulation of the rapid drying process of the coated slurry is essential for achieving the desired structural characteristics. Cachinho et al. [182] described a unique method of preparing a porous Ti scaffold in which the scaffold was prepared by replication of a sponge followed by reactive sintering. The main advantage of this method is the easy production of complex shapes at a low cost. The use of a sacrificial polymeric sponge results in scaffolds with interconnected pores, a critical feature for promoting bone ingrowth and vascularization in newly formed tissue [183]. Such porous architectures are commonly applied in dental implants, permanent osteosynthesis plates, and intervertebral disc replacements [184]. Cachinho et al. [182] reported using 45 vol.% TiH2 powder with a mean particle size of 15.6 µm and a specific surface area of 0.5336 m2/g. The polymeric sponge blocks were dipped into the slurry and infiltrated. After the removal of excess slurry, the samples are dried at room temperature for a period of 24 h. The sintering of samples at a low heating rate of 1 °C/min with dwelling at 500 °C for 2 h and 1000 °C for 4 h results in the formation of a highly porous Ti scaffold with a porosity of 75% and a pore size ranging between 100 µm and 600 µm. The porosity range of 100 µm and 600 µm is appropriate for the growth of new bone tissues and the transport of body fluids. Further, to improve biological properties, the porous Ti is coated with hydroxyapatite and heat-treated at 700 °C. Li et al. [185] synthesized a porous Ti-6Al-4V alloy by replicating 70 wt.% of Ti-6Al-4V powder in water and ammonia solutions. The high-temperature sintering of samples results in open-cell porous Ti struts with a porosity of 88% and a compressive strength of 10 MPa. It should be noticed that the second deposition of powder slurry on the previously sintered scaffold, followed by re-sintering, increases density and compressive strength to 36 MPa. Wang et al. [186], in his study, proposed an improved sponge replication method. A novel solvent consisting of ethanol and water was used to maintain a fast-drying rate and appropriate viscosity of Ti slurry. This slurry was used for multiple Ti coatings, and the Ti scaffold prepared possesses a compressive strength of ~84 MPa with a porosity of ~66%.
Wang et al. [186] explored an enhanced polymeric sponge replication method to fabricate biomedical porous titanium scaffolds. By optimizing key processing parameters, they achieved scaffolds with open and interconnected pore architectures. In their approach, titanium slurries were prepared using a mixture of ethanol and water as the liquid phase. Figure 11a,b display the asymmetrical titanium particles and the open, interconnected structure of the polyurethane (PU) foam used as the sacrificial template. Figure 11c outlines the process flowchart for fabricating porous titanium via this improved replication technique. Notably, three modifications distinguished this method from conventional approaches: (1) Ethanol–water mixtures replaced water alone to improve slurry formulation; (2) excess slurry was removed through centrifugation rather than squeezing, preventing the formation of closed pores; (3) variations in slurry viscosity were introduced for more uniform coating. SEM images in Figure 11d,e illustrate the resulting microstructures, showing that scaffolds fabricated with ethanol exhibited a denser titanium coating compared to those produced using only water. This densification, attributed to ethanol’s anti-foaming properties, resulted in enhanced compressive strength. Furthermore, in vitro studies demonstrated that the scaffolds supported the adhesion, proliferation, and growth of mesenchymal stem cells (MSCs), confirming their suitability for biomedical applications.

6.3. Entangled Metal Wire Technique

Porous Ti implants fabricated via conventional methods exhibit low toughness and tensile strength. The major drawback of these conventional techniques is the difficulty in avoiding contamination and impurities in Ti that arise during processing. Sometimes, undesirable cracks and metallographic defects in sintered Ti struts make them brittle, and thus, they fail to bear tensile load [187]. Also, the porous Ti fabricated by powder metallurgical techniques using a space holder and plasma spray technique has low ductility that may break in the body’s environment when subjected to uncertain overloading and accidents [188]. The entangled metal wire technique (EMWT) is a novel technique that improves these mechanical properties. In this technique, a Ti wire of a diameter of ~0.08 mm to 0.27 mm is used as raw material, and this wire is coiled around a 1.5 mm-diameter rod to form a coiled spring-like structure. The coiled structure is stretched equably such that the distance between two spirals (screw pitch) reaches the external diameter of the coil. This stretched coil is now entangled around a 1 mm diameter rod to form a pre-compacted sample. Finally, this pre-compacted sample is compacted with the help of a piston in a cylindrical die [189].
The mechanism of fabrication and sample porous titanium scaffold prepared by the entangled metal wire technique is shown in Figure 12. Several researchers have fabricated porous titanium scaffolds using EWMT. Zou et al. in 2008 [177] prepared open-cell porous titanium with porosity of 35 to 84% by sintering titanium fibers of 200 µm diameter in a vacuum. The titanium fibers were curved into a helix with the help of a screw, and then, this helix was arranged in a cylindrical form, followed by compaction and vacuum sintering at 1250 °C for 2 h. The resulting porous scaffold has a pore size of 150–600 µm. Young’s modulus was in the range of 3.5–4.2 GPa, and compressive strength was 100–200 MPa. Liu et al. [190] fabricated entangled titanium wire material through different procedures: one with normal wire and another with coiled wire; its compressive and pseudo-elastic hysteresis behavior was investigated. The details of the properties obtained are discussed in Table 5. Jiang et al. [191] fabricated an entangled porous titanium composite filled with biodegradable magnesium melted at 700 °C under the protective environment of SF6 and CO2 to improve the fixation bonding between the implant and host bone. Bisphenol A glycidyl methacrylate (BisGMA) is suggested as a bonding material to provide strong bonding strength and help in fixing the free nodes of the entangled structure [192]. Wang et al. in 2017 [175] proposed a novel technique for the fabrication of a three-dimensional porous titanium scaffold. This method combined two methods: An entangled Molybdenum wire was used as a space holder, and titanium liquid was cast in a vacuum environment, followed by etching off SH particles in an aqua regia solution. The resulting porous scaffold has three-dimensional interconnected pores with porosity in the range of 32–47% and exhibits elastic modulus in the range of 23–62 GPa and yield strength in the range of 76–192 MPa, as shown in Table 5.

6.4. Spark Plasma Sintering (SPS) and Hot Pressing (HP)

The conventional sintering or pressureless sintering process involves heating Ti and its alloys at an elevated temperature of 1200 °C to 1400 °C (but below the melting point of the material) and a high vacuum of the order of 4 × 10−4 Pa to 6 × 10−6 Pa for a long time period of about 24 h to 48 h [194] for densification and homogenization [195,196]. Even after this lengthy procedure, achieving a pore-free homogeneous microstructure is challenging [196]. SPS is an advanced consolidation technique that uses pressure-assisted pulsed current to sinter and can produce porous samples. In this process, the powder is loaded in an electrically conducting die that acts as a heating source when subjected to a pulsed direct current. Thus, the powdered samples will be heated from both sides under uniaxial pressure [197,198,199,200], and due to this fast heating, enhanced mass transfer and rapid powder consolidation will occur [201]. Two theories explain the consolidation mechanism of commercially pure (CP) Ti. According to the first hypothesis, the surface of the powder particles is cleaned and activated by spark discharges generated between metallic powdered particles, thus promoting mass transport for sintering [202,203]. Another hypothesis suggests that the densification of powder is due to particle deformation because, as the temperature increases, the yield strength of the powder particles decreases [204]. Spark plasma sintering (SPS) is also referred to by several alternative names, including field-assisted consolidation, electrical field-activated sintering, plasma-activated sintering, and electrical discharge compaction [203,205,206,207]. SPS is also known by other names, such as field-assisted consolidation technique [205], electrical field-activated sintering [203], plasma-activated sintering [206], and electrical discharge compaction [207]. These methods enable the rapid sintering of metallic powders by applying electrical discharge in combination with swift heating and concurrent pressure. Similarly, in HP, the metallic powder is sintered with the help of electrical resistance in a closed die under uniaxial pressure. Different heating methods, such as induction heating and electric conduction/convection/radiation heating, can be used in HP [208,209]. Ibrahim et al. [210] synthesized porous Ti and its alloy using a cost-effective SPS technique. In this process, porous Ti with different porosities was successfully synthesized by the powder metallurgy technique using NH4HCO3 as a space holder and TiH2 as a foaming agent. SPS is used to consolidate powder at 16 MPa under pressureless conditions. The experimental results showed that pure titanium samples achieved full relative density at a relatively low temperature of 750 °C and a pressure of 16 MPa. The porosity of 53% and Young’s modulus of 40 GPa were achieved in the case of pressureless sintering at a temperature of 1000 °C.
Kashimbetova et al. [211] investigated the fabrication of porous titanium structures using pressure-less spark plasma sintering (PL-SPS). Their study focused on examining how different sintering temperatures influence the microstructural features and mechanical performance of porous Ti. Figure 13a,d,g display the porous strands formed at 1400 °C, where bonding occurred primarily at contact points between adjacent Ti particles in the early sintering stage. At 1500 °C, more developed sintering necks between particles were observed, as shown in Figure 13b,e,h, indicating enhanced densification with limited grain growth. Further sintering at 1600 °C brought the structures into the mid-stage of sintering, marked by significant pore shrinkage, though overall porosity was maintained (Figure 13c,f,i). Increased densification at higher temperatures had a direct effect on compressive yield strength (Figure 13j). At 1400 °C, the weakly bonded particles resulted in a low yield strength of 4.7 MPa. This increased to 26.7 MPa at 1500 °C due to the growth of sintering necks and reached 52.6 MPa at 1600 °C because of further densification. The samples sintered at 1500 °C and 1600 °C exhibited elastic deformation followed by strain hardening during compression. In contrast, the structure sintered at 1400 °C reached peak stress shortly after yielding and then gradually fractured, as shown in Figure 13k. The compressed sample sintered at 1600 °C maintained its pore geometry, particularly in the printing plane (Figure 13l). The findings suggest that intra-strand porosity significantly influences mechanical strength when inter-strand porosity is kept constant. A comparison of the properties of porous titanium synthesized using SPS and hot pressing (HP) from various studies is presented in Table 6.
Figure 13. Images depicting porous titanium fabricated using pressureless spark plasma sintering as a function of varying sintering temperatures: (a,d,g) 1400 °C, (b,e,h) 1500 °C, and (c,f,i) 1600 °C. (j) Compressive stress–strain curves of porous titanium as a function of different sintering temperatures. Images of the (k) fractured sample sintered at 1400 °C and (l) compressed sample (sintered at 1600 °C). Reproduced with permission from [211].
Figure 13. Images depicting porous titanium fabricated using pressureless spark plasma sintering as a function of varying sintering temperatures: (a,d,g) 1400 °C, (b,e,h) 1500 °C, and (c,f,i) 1600 °C. (j) Compressive stress–strain curves of porous titanium as a function of different sintering temperatures. Images of the (k) fractured sample sintered at 1400 °C and (l) compressed sample (sintered at 1600 °C). Reproduced with permission from [211].
Jmmp 09 00401 g013

6.5. Microwave Sintering

Electromagnetic waves with a frequency in the range of 300 MHz to 300 GHz are referred to as microwaves. The most used microwaves for material processing have frequencies between 2.45 GHz and 915 MHz [212]. When materials interact with microwaves, they convert electromagnetic energy into heat energy within the material. The main advantage of microwave sintering is the complete sintering of the material without forming impurities (like oxides). This technique is applicable for sintering ceramics, metals, and composites, offering advantages such as reduced processing time and energy consumption, cost-effectiveness, and environmental sustainability [213]. In powder metallurgy-based manufacturing routes, particularly after pressing and sintering, incompletely bonded or loosely attached particles may remain on the surface of porous titanium structures, and their release into the body can trigger inflammatory responses, macrophage activation, and osteolysis, ultimately jeopardizing implant stability [214]. Preventing such complications requires ensuring adequate particle bonding through optimized sintering conditions and robust neck formation, followed by thorough post-processing steps such as ultrasonic cleaning, chemical/acid etching, abrasive or mechanical removal of weakly attached particles, and thermal treatments to consolidate surface layers. Complementary quality-control procedures, including micro-CT analysis, surface morphology inspection, and mechanical integrity assessments, are also essential to verify the absence of detachable particles. A comprehensive discussion of surface post-treatments for bio-implants is presented in Section 6.6 of this review article. Table 6 demonstrates the processing conditions and physical and mechanical properties obtained using SPS, HP, and microwave sintering.
Table 6. Processing conditions and mechanical properties of porous titanium prepared by spark plasma sintering, hot pressing, and microwave sintering (ST—sintering temperature in °C; TP—time and pressure; P—porosity in %; PS—pore size (µm); E—Young’s modulus (MPa); YS—yield strength in MPa; UCS—ultimate compressive strength in MPa; UTS—ultimate tensile strength in MPa).
Table 6. Processing conditions and mechanical properties of porous titanium prepared by spark plasma sintering, hot pressing, and microwave sintering (ST—sintering temperature in °C; TP—time and pressure; P—porosity in %; PS—pore size (µm); E—Young’s modulus (MPa); YS—yield strength in MPa; UCS—ultimate compressive strength in MPa; UTS—ultimate tensile strength in MPa).
pSTTPPPSEYSUCSUTSRef.
Spark plasma sintering
Pure Ti75016 MPaFully dense-~125---[210]
Pure Ti1000Pressureless53-40---
Ti5Mn alloy950Pressureless56-35---
Ti5Mn alloy1100Pressureless21-52---
Pure Ti700-30–70125–8006–3627–94--[215]
β-alloy Ti-45Nb (gas-atomized)100010 min, 30 MPa0.5 ± 0.1-72 ± 1550--[216]
β-alloy Ti-45Nb (milled)100010 min, 30 MPa4.0 ± 0.2-72 ± 1867--[216]
Ti-6Al-4V7003 min, 30 MPa32 ± 0.2---125-[217]
Pure Ti6003 min, 30 MPa32 ± 0.4---113-[217]
CP Ti (Grade 1) Powder9005 min, 60 MPa---340-445[218,219]
Cryomilled
nanocrystalline CP Ti (Grade 2) powder
850----770-840
CP Ti (Grade 3) powder9005 min, 60 MPa---595-720
Wrought titanium grade 4-3 min, 80 MPa---480–635-655–690
Hot pressing
Ti-45Nb (gas-atomized)60030 min, 700 MPa0.7 ± 0.2-70 ± 1447--[216]
Ti-45Nb (milled)60030 min, 700 MPa3.7 ± 0.1-70 ± 1940--[216]
Microwave sintering
Ti6Al4V/MWCNTi powder1620-25-11 ± 3145270-[213]

6.6. Additive Manufacturing (AM) or Rapid Prototyping (RP)

Additive manufacturing (AM), also known as rapid prototyping (RP), is a computer-assisted advanced fabrication technique that utilizes computer-aided design (CAD) and computer-aided manufacturing (CAM) models to construct predefined microstructures, macrostructures, and precisely controlled hierarchical architectures [220,221]. As the demand of the manufacturing sector is more focused on precision and specific design, the AM technique has emerged as a strong tool for mitigating problems associated with conventional powder metallurgy techniques like material loss during post-processing (like machining, drilling, etc.), large production time, difficulty in producing complex shapes, etc. [222,223,224,225]. AM has been extensively utilized in the medical field by creating patient-specific components based on patients’ medical imaging [226]. The major advantage of this technology is the control over scaffold pore structure, including pore size, shape, volume, and interconnectivity [227]. It is a layer-by-layer fabrication process in which the selected part is built in a CAD file [228,229,230]. The file is sliced along the Z-axis in a virtual environment, and a machine-specific tool path is generated for each slice.
AM offers specific advantages like site-specific deposition with higher cooling rates and can easily produce intricate shapes [231]. The interconnectivity of the porous structures is a key issue in porous metallic structures. As mentioned above, this issue cannot be perfectly controlled by other conventional manufacturing techniques. In contrast, the final structure of the AM-fabricated components relies on the initial design with interconnected structures through AM. The porous Ti can be easily fabricated without design restrictions. Additive manufacturing (AM) techniques can be categorized based on the energy–powder interaction and consolidation mechanism, including methods such as electron beam melting (EBM), selective laser melting (SLM)—also referred to as laser powder bed fusion (LPBF)—laser-engineered net shaping (LENS), and binder jetting (BJG) [232,233]. Though there are several AM processes, only a few techniques, including SLM/LPBF and EBM, are widely used and can fabricate a wide variety of materials. Other AM processes like directed energy deposition (DED) are widely used for higher deposition rates and large-scale manufacturing [234].
Apart from fabricating intricate shapes, the AM structures also exhibit improved properties because of hierarchical microstructures resulting from higher cooling rates observed during the process [235,236,237]. Moreover, AM produces a unique microstructure and texture compared to conventional processing, which helps to improve the strength by maintaining the same porosity [238]. As mentioned in Section 5, Ti shows allotropy. Moreover, Ti6Al4V exhibits the coexistence of α and β microstructures, providing versatility in the Ti-based alloys. Since SLM/LPBF offers high cooling rates, a fine α’ martensite or α’ martensite-based microstructure is observed in these samples [239,240,241]. The α’ martensite appears as an acicular or needle-shaped microstructure, which has an hcp crystal structure like the α phase. Like ferrous martensite, Ti-based martensite also helps improve its strength at the cost of ductility. Hence, AM-fabricated structures may require post-processing, such as heat treatment, to tune their properties according to the requirements [242,243]. Sabban et al. [244] introduced a novel heat treatment approach that successfully transformed martensitic laths into a bimodal globularized microstructure in the AM Ti64 alloy, resulting in an 80% enhancement of ductility and a 66% improvement in toughness. In subsequent research by the same group, Gupta et al. [245] fabricated a Ti64-based bone plate using the AM technique. They employed a specially designed process of repeated cyclic heating and cooling, as previously described, to transform the microstructure from acicular to bimodal without the need for plastic deformation before heat treatment. Moreover, the in situ dissolution of α’ is also possible by changing the processing parameters and providing double or triple laser scanning in selective laser manufacturing [246,247,248].
AM-fabricated porous Ti alloys exhibit relatively better mechanical properties due to interconnected structures [35,249]. The good interconnectivity of the porous implants enables uniform load distribution among the structures, while poorly connected porous structures typically fail prematurely at the joints. Ample evidence in the literature suggests premature failure for porous metallic implants [250,251]. Given the advantages, there are also some drawbacks, including surface defects, which usually tend to deteriorate the build’s properties. One can understand the importance of AM for biomedical applications, as illustrated in Figure 14, Figure 15, Figure 16, Figure 17 and Figure 18. Biomedical implant dimensions and requirements vary from patient to patient, making other routes challenging. Porous Ti implants are usually used for orthopedic implants owing to their outstanding mechanical and biological properties. The elastic modulus of bone ranges between 10 and 20 GPa and is mostly composed of inner cancellous and outer cortical bone [252]. Meanwhile, dense Ti alloys have an elastic modulus of around 110 GPa. However, it is worth mentioning that metallic alloys fabricated via AM tend to possess a lower elastic modulus than regular Ti-based alloys, making them comparable to bone [32]. In addition, porous Ti also possesses high rigidity and good fatigue properties, making it a suitable candidate for biomedical implants.
Figure 14. (a) Medical titanium alloy for anti-corrosive properties in femoral implant [253]; (b) 3D-printed titanium hip ball-and-socket implant: image credit: Jabil [254]; (c) first trabecular titanium hip cup produced by Lima Corporate, San Daniele del Friuli, Italy and Arcam Mölndal, Sweden (now a GE Additive company) in the year 2007 [255].
Figure 14. (a) Medical titanium alloy for anti-corrosive properties in femoral implant [253]; (b) 3D-printed titanium hip ball-and-socket implant: image credit: Jabil [254]; (c) first trabecular titanium hip cup produced by Lima Corporate, San Daniele del Friuli, Italy and Arcam Mölndal, Sweden (now a GE Additive company) in the year 2007 [255].
Jmmp 09 00401 g014
Figure 14 shows the Ti femoral implant and the hip ball-and-socket joints [253,254,255]. Pelvic injury is prevalent in trauma patients, caused by impact, rolling, steep falls, and other injuries. Recently, AM has often been employed for preoperative examination and simulated surgery treating pelvic fractures. Investigations demonstrate the possibility of employing AM-fabricated Ti alloys for pelvic implants [256,257]. Wong et al. [258] manufactured pelvic-specific implants from Ti alloy and evaluated their performance. Broekhuis et al. [259] used AM to customize and create metallic pelvic prostheses for acetabular repair following tumor resections.
The spinal column is composed of interconnecting bones called vertebrae, which begin at the base of the head and finish at the tailbone in the lower back. The vertebral column is formed by stacking one bony vertebra on top of the next. The intervertebral discs are disc cushions that sit between each of the vertebral bodies. These discs act as shock absorbers, connecting the vertebrae (Figure 15). These discs also allow for the spine’s bending and twisting action, assisted by facet joints. The spinal column protects the spinal cord, which links the brain’s nerves to other body parts. In 2017, the United States FDA (Food and Drug Administration) authorized two alloys of Ti spinal implants made by AM. One example of an additively manufactured vertebral implant is the HAWKEYE Ti, while another is the NEXXT MATRIXX 3D-printed spinal implant. Hollander et al. [260,261] employed AM technology to fabricate spinal implants with tailored porosity and demonstrated that their surface characteristics supported the adhesion and proliferation of human osteoblasts. Lin et al. [261] utilized the SLM/LPBF method to achieve a porosity of approximately 55%, resulting in an elastic modulus close to that of natural human bone (~3 GPa). This structure also exhibited superior bone regeneration capabilities compared to conventional PEEK cages.
Figure 15. Human spinal column with the insert in black illustrating ATEC’s IdentiTi posterior curved porous titanium interbody implants [262].
Figure 15. Human spinal column with the insert in black illustrating ATEC’s IdentiTi posterior curved porous titanium interbody implants [262].
Jmmp 09 00401 g015
In certain cases, implants can act as bone replacements without being subjected to considerable and constant bearing or additional stress. Figure 16 is the cranial implant developed by Novax DMA [263] for a patient requiring a large metallic implant. In the case of a skull implant, external variables, particularly the implant itself, should be promoted rather than hindering the healing process. Most significantly, the implant should fit as perfectly as feasible, as offered by AM. The manufacturing technique, which uses a laser for building up a material (Ti) layer by layer, allows for maximum customization in terms of form and size. Because of the large hole in the patient’s bone structure, integrating biological activities and minimizing heat loss into the brain tissue were given high priority. A Ti construction may also be impermeable to brain tissue fluid. Only permeable construction may satisfy the requirements.
Figure 16. (a) Skull implants developed by Novax DMA. (b) Magnified image of the porous titanium cranial implants with porosity levels of 95% [263].
Figure 16. (a) Skull implants developed by Novax DMA. (b) Magnified image of the porous titanium cranial implants with porosity levels of 95% [263].
Jmmp 09 00401 g016
A lattice-structured implant with skull-integrated, screw-in fasteners allows fluids to pass through and combine with the skull’s bone. Furthermore, such a construction would provide insulation, reducing heat transfer entering the cranial cavity. The pore spaces are around 1 mm, whereas the cell links are about 0.2 mm thick. Murr et al. [264] described a reticular skull implant developed through EBM. Mazzoli et al. [265] used CT imaging, computer modeling, and AM to produce a biocompatible Ti skull implant. Zhao et al. [266] observed that personalized Ti-based skull prosthetics made via AM offered higher impact resistance while efficiently repairing skull deformities and protecting intracranial brain tissue. Yan et al. [267] employed EBM to create a mandibular prosthetic implant that had a 3D mesh and a porosity of 81.3%, which fits the parameters for implantation in the human body (Figure 17). Figure 17a illustrates the 3D anatomic model of the human mandible developed using a CT imaging technique. Figure 17b shows the 3D structure developed in CAD software (Unigraphics NX 8.0, EDS) by providing internal 3D meshing for porous structures. Meanwhile, Figure 17c,d demonstrate the mandibular prosthetic implant produced by EBM. Moiduddin et al. [268] utilized a Ti-based alloy to fabricate cheekbones.
Figure 17. (a) Three-dimensional anatomic model of a human mandible prosthetic. (b) Design of mandibular scaffold 3D mesh internal microstructure. (c,d) Titanium alloy scaffolds produced by EBM with porosity levels of 81.3% and a weight of 107 g. Reproduced with permissions from [267].
Figure 17. (a) Three-dimensional anatomic model of a human mandible prosthetic. (b) Design of mandibular scaffold 3D mesh internal microstructure. (c,d) Titanium alloy scaffolds produced by EBM with porosity levels of 81.3% and a weight of 107 g. Reproduced with permissions from [267].
Jmmp 09 00401 g017
Yanez et al. [269] studied the gyroid porous Ti structures, where triply periodic minimum surfaces (TPMS) have been recognized as an effective approach for developing porous biomaterials. They developed a variety of porous Ti structures with varying degrees of porosities using EBM to investigate mechanical characteristics under compression and torsion loads, as illustrated in Figure 18. Compression testing indicated that deformed gyroid structures possess significant strength and stiffness when subjected to axial loading, particularly when a reinforcing shell is incorporated. In torsional loading, however, conventional gyroids outperformed deformed gyroids with high CAD-defined porosity (90%) in terms of both torsional strength and stiffness. Notably, deformed gyroids exhibited superior mechanical behavior under compressive loads compared to their regular counterparts. Conversely, regular gyroids demonstrated better adaptability to non-axial and torsional loading conditions [269].
Figure 18. (a) Six variations of gyroid-structured porous titanium: normal gyroid 75, normal gyroid 90, deformed gyroid 75, deformed gyroid 90, deformed gyroid with shell reinforcement, and double deformed gyroid. (b) Visual depiction of compressive testing performed on the gyroid porous titanium structures. (c) Image capturing the torsional testing of one of the porous titanium configurations. Reproduced with permissions from [269].
Figure 18. (a) Six variations of gyroid-structured porous titanium: normal gyroid 75, normal gyroid 90, deformed gyroid 75, deformed gyroid 90, deformed gyroid with shell reinforcement, and double deformed gyroid. (b) Visual depiction of compressive testing performed on the gyroid porous titanium structures. (c) Image capturing the torsional testing of one of the porous titanium configurations. Reproduced with permissions from [269].
Jmmp 09 00401 g018
Figure 19 depicts the many surface flaws that arise during bio-implant manufacturing. AM-processed bio-implants are often not ready for use due to surface defects such as powder adhesions, semi-welds, surface casing, porosity, balling effect, etc., highlighting the need for post-processing. The nature of the designs, prior processing software, and processing equipment all influence post-processing requirements. Post-processing includes removing unwanted material, improving the texture of the surface, aesthetic enhancements, component separation, rebinding and sintering, machining, surface finishing, non-thermal performance enhancement, heat treatment, and quality assurance. It is critical to enhance the surface finish by employing suitable surface quality techniques for AM-built biomedical implants to compete alongside widely recognized traditionally fabricated bio-implants [270,271,272].
Figure 19. Surface defects occurring during AM-fabricated implants (reproduced with permission from [272]).
Figure 19. Surface defects occurring during AM-fabricated implants (reproduced with permission from [272]).
Jmmp 09 00401 g019
Conventional processes involve employing multipoint or single-point instruments for cutting directly over the surface that is being finished or polished. These include robot-based finishing, rigid tool-based finishing, and computer numerical control-based finishing. Conventional processes are more difficult and time-consuming as they tend to induce residual stress on the surface. Cutting speed, feed, and other factors impact the machining characteristics and efficacy of difficult-to-cut minerals or metals, particularly Ti. The right choice of cutting and finishing parameters is critical for increased production and widespread use in industries. Because of the increased use of implant finishing and existing issues with traditional machining, innovative methods are necessary to improve the machining properties of metallic implants [273]. The finishing procedure comprises polishing, but only for the outside cylindrical sections; it cannot complete the inside cylinders or intricate geometrical forms of the implants. Similarly, the grinding method of finishing does not allow access to the implants’ deep holes and may leave markings on the final surface. Lapping is a laborious finishing method that is not suitable for intricate geometrical designs of implants. The burnishing procedure causes the completed surface to harden, making it unsuitable for thin-walled surfaces [274]. All limitations of conventional procedures have necessitated the development of improved finishing methods for bioimplants. Figure 20 illustrates the categories of surface post-treatments that may improve the surface characteristics of traditional or additively made bio-implants.
Figure 20. Classification of surface post-treatment for titanium implant materials for enhanced surface properties. Reproduced with permission from [270,271,272].
Figure 20. Classification of surface post-treatment for titanium implant materials for enhanced surface properties. Reproduced with permission from [270,271,272].
Jmmp 09 00401 g020
Srijan et al. [275] utilized surface mechanical attrition treatment (SMAT), a severe plastic deformation technique, to enhance the surface-dependent and functional properties of β-Ti alloy. The increased defect density resulting from SMAT led to a reduced corrosion rate through the formation of a stronger oxide layer and an improvement in wear resistance. Srijan et al. reviewed and classified recent advancements in surface modification, focusing on the induction of nano-crystallization on metallic biomaterials through various surface modification techniques [276].

6.7. Recent Trends in the Development of Porous Ti Scaffold: Titanium-Based Interpenetrating Phase Composites

As discussed in the previous sections, the major causes of implant failure, like stress shielding, implant loosening, etc., can be mitigated by reducing Young’s modulus by making them porous using different fabrication techniques. However, these modifications invariably reduce the strength and fatigue resistance as compared to their dense counterpart [31,277,278,279,280]. To enhance the mechanical and biological properties of Ti-based scaffolds, researchers explore various techniques to integrate a secondary bioactive phase into the 3D-printed porous framework [281]. Magnesium, zinc, calcium, and iron [44,282,283,284,285], or their alloys, are commonly utilized metallic biomaterials. At the same time, hydroxyapatite (HAp), wollastonite (CaSiO3) [28], etc., are common ceramic materials that can serve as filler material. The logic behind adding all these materials inside the metallic porous network is to provide mechanical and biological support during the initial implantation period. Since these filler materials are either biodegradable or bioresorbable, they naturally degrade within the human body over time [286]. Zhang et al. [287] developed a Mg-Ti composite via the pressureless infiltration of Mg into an AM-fabricated Ti6Al4V scaffold. Since the constituents of the composite were continuous and mutually interpenetrated in the 3-D space, the composite was termed an interpenetrating phase composite. These porous structures were designed to mimic the bioinspired architecture (Figure 21). The composite shows effective stress transfer, delocalized damage, and arrest cracking. Similarly, Dou et al. [281] fabricated partially degradable interpenetrating phase composites by first creating a pure titanium-based dodecahedral cellular structure with pore sizes ranging from 400 to 500 μm using selective laser melting (SLM), followed by pressureless infiltration of pure magnesium into the porous framework. The developed composite possesses a higher strength (yield strength of ~64 MPa and ultimate compressive strength of ~275 MPa) and lower modulus (~47.3 GPa) than cast pure Mg and Ti. Also, the degradation of the Mg supports the ingrowth of bone tissue and biological fixation between the host and implant. In vitro and in vivo studies of these composites revealed that the composite exhibits accelerated corrosion compared to pure Mg. Still, it remains non-cytotoxic and does not induce obvious adverse reactions following implantation.
In a similar approach, Rahmani et al. [28] reported the addition of three different types of wollastonite (CaSiO3)-based bio-ceramic via SPS in an AM Ti6Al4V lattice, as shown in Figure 22. The produced composite exhibited higher wear resistance, damage tolerance, and mechanical properties, which can improve the durability of the bones. In addition to Ti6Al4V, TiNi, and Ti22Al25Nb, porous samples infiltrated with wollastonite showed that TiNi has a higher damping capacity, while Ti22Al25Nb induces a higher portion of elastic deformation. In another study, Rahmani et al. [28] reported improved impact and osteoinductivity by manufacturing hybrid wollastonite and Ti6Al4V for craniofacial implants.

7. Current Challenges for Porous Titanium

Porous Ti has the potential to be suited for biomedical applications due to its high strength, lower elastic modulus, longer fatigue life, and good wear and corrosion resistance. In addition to the abovementioned mechanical properties, porous Ti exhibits excellent biocompatibility and helps pass fluids through the porous structures. Nevertheless, challenges are associated with the fabrication and processing of porous Ti implants:
(1)
Porous Ti implants are difficult to manufacture, and optimally controlling pores and maintaining the uniformity of pores are challenging tasks.
(2)
The interconnectivity of the pores is essential in determining the mechanical properties of the implants. It is paramount to maintain a tradeoff between the strength and porosity of porous Ti implants.
(3)
As implants undergo repeated cycles of loading and unloading during daily activities, the porous structures generally exhibit lower fatigue resistance compared to their dense counterpart. The pores serve as potential initiation points for fatigue cracks, which can gradually propagate and ultimately result in premature implant failure.
(4)
There must be a balance between patient-specific implants and large-scale production of implants, i.e., customization and production, since developing porous structures and meshing is time-consuming.
(5)
The implant cost should also be considered, as porous Ti implants are usually costlier than fully dense implants. Higher costs of implants reduce the demand in the market.
(6)
Implementing thorough checking to minimize defect concentration is important to obtain high-quality implants with better mechanical properties and biocompatibility.
(7)
Especially in the case of AM, thermal gradients can lead to variations in pore sizes and porosity levels.
(8)
The post-processing of the porous structures is also very difficult and time-consuming. It must be carried out carefully, as small disruptions can damage the interconnected structures, leading to defective implants.

8. Future Scope of Titanium-Based Porous Implants

Porous Ti has been proven to be the best-suited candidate, especially for orthopedic implants and prosthetics. It has taken the field of biomaterials, specifically implants, by storm and is expected to be the future of biomedical implants, especially orthopedic implants. However, in short, a few porous Ti challenges must be resolved to make it irreplaceable in the market:
(1)
AM has been gaining momentum in producing porous Ti structures. However, more efforts must be taken to fabricate porous structures with good pore interconnectivity, which will, in turn, pave the way for better mechanical properties and biocompatibility.
(2)
A biocompatible coating can be applied to enhance the osseointegration of the implant with body fluids.
(3)
Porous Ti structures can be used as multifunctional implants by integrating them with drug delivery systems or sensors.
(4)
Efforts should be made to make the porous Ti implants more accessible at an affordable price, which is possible through process optimization to achieve higher productivity at a lower cost.
(5)
Investigations of biocompatible joining strategies should be carried out for integrating porous Ti into various biomaterials or metallic implant structures.
(6)
The development of interpenetrating phase composites is limited to the Ti-Mg system since the developed composite has the potential to be utilized as an implant; therefore, a composite system comprising Ti-Zn/Ca/Fe and SS-Mg/Zn/Ca/Fe should be explored.

9. Conclusions

The present review focused on exploring the development of Ti-based porous biomaterial for orthopedic applications. Different characteristics and requirements of biomaterials, like mechanical, physical, and biological properties, are comprehensively reviewed. A section of the review was dedicated to the structure of bone and its physiology for the basic understanding of bone metabolism, which should be considered while designing and synthesizing next-generation orthopedic and dental implants. Considering the importance of porous Ti-based implants for orthopedic applications, this review presents different processing techniques, particularly based on powder metallurgy processes, and the properties of the developed porous implants are compared. Given the importance of the powder metallurgical route for synthesizing porous implants, a detailed study of various powder processing techniques was carried out. Finally, this review concludes by discussing the current challenges and future potential of porous Ti-based alloys for biomaterial implant application.

Author Contributions

Conceptualization, M.K.Y., P.K., J.J. and P.K.G.; methodology, M.K.Y., A.Y., Y.N.A., P.K. and V.P.; validation, C.S.P., A.N., K.C., S.S., J.J. and P.K.G.; formal analysis, M.K.Y. and A.Y.; investigation, M.K.Y., A.Y., P.K., V.P. and C.S.P.; resources, S.S., K.C. and P.K.G.; data curation, M.K.Y., A.Y. and P.K.; writing—original draft preparation, M.K.Y., A.Y., Y.N.A. and P.K.; writing—review and editing, V.P., C.S.P., A.N., K.C., S.S., J.J. and P.K.G.; supervision, P.K.G.; project administration, P.K.G. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the European Union through the REFRESH—Research Excellence for Region Sustainability and High-tech Industries—project (Project No. CZ.10.03.01/00/22_003/0000048), funded under the Operational Programme Just Transition.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Cui, Y.W.; Wang, L.; Zhang, L.C. Towards Load-Bearing Biomedical Titanium-Based Alloys: From Essential Requirements to Future Developments. Prog. Mater. Sci. 2024, 144, 101277. [Google Scholar] [CrossRef]
  2. Banhart, J. Manufacture, Characterisation and Application of Cellular Metals and Metal Foams. Prog. Mater. Sci. 2001, 46, 559–632. [Google Scholar] [CrossRef]
  3. Dabrowski, B.; Swieszkowski, W.; Godlinski, D.; Kurzydlowski, K.J. Highly Porous Titanium Scaffolds for Orthopaedic Applications. J. Biomed. Mater. Res. B Appl. Biomater. 2010, 95, 53–61. [Google Scholar] [CrossRef] [PubMed]
  4. Nabiyouni, M.; Brückner, T.; Zhou, H.; Gbureck, U.; Bhaduri, S.B. Magnesium-Based Bioceramics in Orthopedic Applications. Acta Biomater. 2018, 66, 23–43. [Google Scholar] [CrossRef]
  5. Kim, T.; See, C.W.; Li, X.; Zhu, D. Orthopedic Implants and Devices for Bone Fractures and Defects: Past, Present and Perspective. Eng. Regen. 2020, 1, 6–18. [Google Scholar] [CrossRef]
  6. Spicer, P.P.; Kretlow, J.D.; Young, S.; Jansen, J.A.; Kasper, F.K.; Mikos, A.G. Evaluation of Bone Regeneration Using the Rat Critical Size Calvarial Defect. Nat. Protoc. 2012, 7, 1918–1929. [Google Scholar] [CrossRef]
  7. Oryan, A.; Alidadi, S.; Moshiri, A.; Maffulli, N. Bone Regenerative Medicine: Classic Options, Novel Strategies, and Future Directions. J. Orthop. Surg. Res. 2014, 9, 18. [Google Scholar] [CrossRef]
  8. Zhou, H.; Lawrence, J.G.; Bhaduri, S.B. Fabrication Aspects of PLA-CaP/PLGA-CaP Composites for Orthopedic Applications: A Review. Acta Biomater. 2012, 8, 1999–2016. [Google Scholar] [CrossRef]
  9. Stevens, M.M. Biomaterials for Bone Tissue Engineering. Mater. Today 2008, 11, 18–25. [Google Scholar] [CrossRef]
  10. Keating, J.F.; McQueen, M.M. Substitutes for autologous bone graft in orthopaedic trauma. J. Bone Jt. Surg. Br. Vol. 2001, 83, 3–8. [Google Scholar] [CrossRef]
  11. Damien, C.J.; Parsons, J.R. Bone Graft and Bone Graft Substitutes: A Review of Current Technology and Applications. J. Appl. Biomater. 1991, 2, 187–208. [Google Scholar] [CrossRef] [PubMed]
  12. Hering, B.J.; Walawalkar, N. Pig-to-Nonhuman Primate Islet Xenotransplantation. Transpl. Immunol. 2009, 21, 81–86. [Google Scholar] [CrossRef] [PubMed]
  13. Fishman, J.A.; Patience, C. Xenotransplantation: Infectious Risk Revisited. Am. J. Transplant. 2004, 4, 1383–1390. [Google Scholar] [CrossRef] [PubMed]
  14. Cooper, D.K.C.; Ezzelarab, M.B.; Hara, H.; Iwase, H.; Lee, W.; Wijkstrom, M.; Bottino, R. The Pathobiology of Pig-to-Primate Xenotransplantation: A Historical Review. Xenotransplantation 2016, 23, 83–105. [Google Scholar] [CrossRef]
  15. 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]
  16. Yadav, M.K.; Pandey, V.; Jyoti; Kumar, A.; Mohanta, K.; Singh, V.K. Mechanical and Biological Behaviour of Porous Ti–SiO2 Scaffold for Tissue Engineering Application. Ceram. Int. 2021, 47, 22191–22200. [Google Scholar] [CrossRef]
  17. Kumar Yadav, M.; Hiren Shukla, R.; Prashanth, K.G. A Comprehensive Review on Development of Waste Derived Hydroxyapatite (HAp) for Tissue Engineering Application. Mater. Today Proc. 2023. [Google Scholar] [CrossRef]
  18. Yadav, M.K.; Pandey, V.; Mohanta, K.; Singh, V.K. A Low-Cost Approach to Develop Silica Doped Tricalcium Phosphate (TCP) Scaffold by Valorizing Animal Bone Waste and Rice Husk for Tissue Engineering Applications. Ceram. Int. 2022, 48, 25335–25345. [Google Scholar] [CrossRef]
  19. Ershad-Langroudi, A.; Babazadeh, N.; Alizadegan, F.; Mehdi Mousaei, S.; Moradi, G. Polymers for Implantable Devices. J. Ind. Eng. Chem. 2024, 137, 61–86. [Google Scholar] [CrossRef]
  20. Singh, N.; Srikanth, K.P.; Gopal, V.; Rajput, M.; Manivasagam, G.; Prashanth, K.G.; Chatterjee, K.; Suwas, S. In Situ Production of Low-Modulus Ti–Nb Alloys by Selective Laser Melting and Their Functional Assessment toward Orthopedic Applications. J. Mater. Chem. B 2024, 12, 5982–5993. [Google Scholar] [CrossRef]
  21. Ummethala, R.; Karamched, P.S.; Rathinavelu, S.; Singh, N.; Aggarwal, A.; Sun, K.; Ivanov, E.; Kollo, L.; Okulov, I.; Eckert, J.; et al. Selective Laser Melting of High-Strength, Low-Modulus Ti–35Nb–7Zr–5Ta Alloy. Materialia 2020, 14, 100941. [Google Scholar] [CrossRef]
  22. Hameed, P.; Liu, C.F.; Ummethala, R.; Singh, N.; Huang, H.H.; Manivasagam, G.; Prashanth, K.G. Biomorphic Porous Ti6Al4V Gyroid Scaffolds for Bone Implant Applications Fabricated by Selective Laser Melting. Prog. Addit. Manuf. 2021, 6, 455–469. [Google Scholar] [CrossRef]
  23. Khan, P.A.; Raheem, A.; Kalirajan, C.; Prashanth, K.G.; Manivasagam, G. In Vivo Assessment of a Triple Periodic Minimal Surface Based Biomimmetic Gyroid as an Implant Material in a Rabbit Tibia Model. ACS Mater. Au 2024, 4, 479–488. [Google Scholar] [CrossRef] [PubMed]
  24. Neto, J.V.C.; Teixeira, A.B.V.; Cândido dos Reis, A. Hydroxyapatite Coatings versus Osseointegration in Dental Implants: A Systematic Review. J. Prosthet. Dent. 2023, 134, 92–99. [Google Scholar] [CrossRef] [PubMed]
  25. Salahinejad, E.; Vahedifard, R. Deposition of Nanodiopside Coatings on Metallic Biomaterials to Stimulate Apatite-Forming Ability. Mater. Des. 2017, 123, 120–127. [Google Scholar] [CrossRef]
  26. Rahmani, R.; Lopes, S.I.; Prashanth, K.G. Selective Laser Melting and Spark Plasma Sintering: A Perspective on Functional Biomaterials. J. Funct. Biomater. 2023, 14, 521. [Google Scholar] [CrossRef]
  27. Rahmani, R.; Kamboj, N.; Brojan, M.; Antonov, M.; Prashanth, K.G. Hybrid Metal-Ceramic Biomaterials Fabricated through Powder Bed Fusion and Powder Metallurgy for Improved Impact Resistance of Craniofacial Implants. Materialia 2022, 24, 101465. [Google Scholar] [CrossRef]
  28. Rahmani, R.; Antonov, M.; Kollo, L.; Holovenko, Y.; Prashanth, K.G. Mechanical Behavior of Ti6Al4V Scaffolds Filled with CaSiO3 for Implant Applications. Appl. Sci. 2019, 9, 3844. [Google Scholar] [CrossRef]
  29. Kamboj, N.; Rodríguez, M.A.; Rahmani, R.; Prashanth, K.G.; Hussainova, I. Bioceramic Scaffolds by Additive Manufacturing for Controlled Delivery of the Antibiotic Vancomycin. Proc. Est. Acad. Sci. 2019, 68, 185–190. [Google Scholar] [CrossRef]
  30. Al-Shalawi, F.D.; Azmah Hanim, M.A.; Ariffin, M.K.A.; Looi Seng Kim, C.; Brabazon, D.; Calin, R.; Al-Osaimi, M.O. Biodegradable Synthetic Polymer in Orthopaedic Application: A Review. Mater. Today Proc. 2023, 74, 540–546. [Google Scholar] [CrossRef]
  31. Praveenkumar, K.; Manivasagam, G.; Swaroop, S. Effect of Laser Peening on the Residual Stress Distribution and Wettability Characteristics of Ti-6Al-4V Alloy for Biomedical Applications. Trends Biomater. Artif. Organs 2022, 36, 18–25. [Google Scholar]
  32. Arabnejad, S.; Johnston, B.; Tanzer, M.; Pasini, D. Fully Porous 3D Printed Titanium Femoral Stem to Reduce Stress-Shielding Following Total Hip Arthroplasty. J. Orthop. Res. 2017, 35, 1774–1783. [Google Scholar] [CrossRef] [PubMed]
  33. Engler, I.D.; Hart, P.A.; Swanson, D.P.; Kirsch, J.M.; Murphy, J.P.; Wright, M.A.; Murthi, A.; Jawa, A. High Prevalence of Early Stress Shielding in Stemless Shoulder Arthroplasty. Semin. Arthroplast. JSES 2022, 32, 751–756. [Google Scholar] [CrossRef]
  34. Nagels, J.; Stokdijk, M.; Rozing, P.M. Stress Shielding and Bone Resorption in Shoulder Arthroplasty. J. Shoulder Elb. Surg. 2003, 12, 35–39. [Google Scholar] [CrossRef]
  35. Praveenkumar, K.; Vishnu, J.; Samuel S., C.; Gopal, V.; Arivarasu, M.; Lackner, J.M.; Meier, B.; Karthik, D.; Suwas, S.; Swaroop, S.; et al. High Temperature Dry Sliding Wear Behaviour of Selective Laser Melted Ti-6Al-4V Alloy Surfaces. J. Mater. Process. Technol. 2024, 329, 118439. [Google Scholar] [CrossRef]
  36. Praveenkumar, K.; Swaroop, S.; Manivasagam, G. Effect of Multiple Laser Shock Peening without Coating on Residual Stress Distribution and High Temperature Dry Sliding Wear Behaviour of Ti-6Al-4 V Alloy. Opt. Laser Technol. 2023, 164, 109398. [Google Scholar] [CrossRef]
  37. Praveenkumar, K.; Mylavarapu, P.; Sarkar, A.; Isaac Samuel, E.; Nagesha, A.; Swaroop, S. Residual Stress Distribution and Elevated Temperature Fatigue Behaviour of Laser Peened Ti-6Al-4V with a Curved Surface. Int. J. Fatigue 2022, 156, 106641. [Google Scholar] [CrossRef]
  38. Praveenkumar, K.; Vishnu, J.; Raheem, A.; Gopal, V.; Swaroop, S.; Suwas, S.; Shankar, B.; Manivasagam, G. In-Vitro Fretting Tribocorrosion and Biocompatibility Aspects of Laser Shock Peened Ti-6Al-4V Surfaces. Appl. Surf. Sci. 2024, 665, 160334. [Google Scholar] [CrossRef]
  39. Bandyopadhyay, A.; Mitra, I.; Goodman, S.B.; Kumar, M.; Bose, S. Improving Biocompatibility for next Generation of Metallic Implants. Prog. Mater. Sci. 2023, 133, 101053. [Google Scholar] [CrossRef]
  40. Bai, L.; Gong, C.; Chen, X.; Sun, Y.; Zhang, J.; Cai, L.; Zhu, S.; Xie, S.Q. Additive Manufacturing of Customized Metallic Orthopedic Implants: Materials, Structures, and Surface Modifications. Metals 2019, 9, 1004. [Google Scholar] [CrossRef]
  41. Donachie, M.J. Titanium: A Technical Guide; ASM International: Almere, The Netherlands, 2000. [Google Scholar] [CrossRef]
  42. Sanan, A.; Haines, S.J. Repairing Holes in the Head: A History of Cranioplasty. Neurosurgery 1997, 40, 588–603. [Google Scholar] [CrossRef] [PubMed]
  43. Praveenkumar, K.; Mylavarapu, P.; Swaroop, S. Surface Oxidation and Subsurface Deformation in a Laser-Peened Ti-6Al-4V. J. Mater. Eng. Perform. 2023, 32, 7348–7362. [Google Scholar] [CrossRef]
  44. Vishnu, J.; Praveenkumar, K.; Kumar, A.A.; Nair, A.; Arjun, R.; Pillai, V.G.; Shankar, B.; Shankar, K.V. Multifunctional Zinc Oxide Loaded Stearic Acid Surfaces on Biodegradable Magnesium WE43 Alloy with Hydrophobic, Self-Cleaning and Biocompatible Attributes. Appl. Surf. Sci. 2025, 680, 161455. [Google Scholar] [CrossRef]
  45. 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]
  46. Müller, U.; Imwinkelried, T.; Horst, M.; Sievers, M.; Graf-Hausner, U. Do Human Osteoblasts Grow into Open-Porous Titanium? Cell Mater 2006, 11, 8–15. [Google Scholar] [CrossRef]
  47. Pohler, O.E.M. Unalloyed Titanium for Implants in Bone Surgery. Injury 2000, 31, D7–D13. [Google Scholar] [CrossRef]
  48. Yang, S.; Leong, K.F.; Du, Z.; Chua, C.K. The Design of Scaffolds for Use in Tissue Engineering. Part I. Traditional Factors. Tissue Eng. 2001, 7, 679–689. [Google Scholar] [CrossRef]
  49. Maity, T.; Balcı, Ö.; Gammer, C.; Ivanov, E.; Eckert, J.; Prashanth, K.G. High Pressure Torsion Induced Lowering of Young’s Modulus in High Strength TNZT Alloy for Bio-Implant Applications. J. Mech. Behav. Biomed. Mater. 2020, 108, 103839. [Google Scholar] [CrossRef]
  50. Nag, S.; Samuel, S.; Puthucode, A.; Banerjee, R. Characterization of Novel Borides in Ti-Nb-Zr-Ta + 2B Metal-Matrix Composites. Mater. Charact. 2009, 60, 106–113. [Google Scholar] [CrossRef]
  51. Luo, J.P.; Huang, Y.J.; Xu, J.Y.; Sun, J.F.; Dargusch, M.S.; Hou, C.H.; Ren, L.; Wang, R.Z.; Ebel, T.; Yan, M. Additively Manufactured Biomedical Ti-Nb-Ta-Zr Lattices with Tunable Young’s Modulus: Mechanical Property, Biocompatibility, and Proteomics Analysis. Mater. Sci. Eng. C 2020, 114, 110903. [Google Scholar] [CrossRef]
  52. Sahay, S.; Kesavan, P.; Yadav, M.K.; Nilawar, S.; Manivasagam, G.; Chatterjee, K. High-Pressure Torsion Affects Mechanical Properties, Electrochemical Behavior, and Cellular Response to a Biomedical Ti-Nb-Zr-Ta Alloy. Mater. Trans. 2025, 66, 490–500. [Google Scholar] [CrossRef]
  53. Weber, J.N.; White, E.W. Carbon-Metal Graded Composites for Permanent Osseous Attachment of Non-Porous Metals. Mater. Res. Bull. 1972, 7, 1005–1016. [Google Scholar] [CrossRef]
  54. Prashanth, K.; Zhuravleva, K.; Okulov, I.; Calin, M.; Eckert, J.; Gebert, A. Mechanical and Corrosion Behavior of New Generation Ti-45Nb Porous Alloys Implant Devices. Technologies 2016, 4, 33. [Google Scholar] [CrossRef]
  55. Zhang, L.; He, Z.Y.; Tan, J.; Calin, M.; Prashanth, K.G.; Sarac, B.; Völker, B.; Jiang, Y.H.; Zhou, R.; Eckert, J. Designing a Multifunctional Ti-2Cu-4Ca Porous Biomaterial with Favorable Mechanical Properties and High Bioactivity. J. Alloys Compd. 2017, 727, 338–345. [Google Scholar] [CrossRef]
  56. Attar, H.; Löber, L.; Funk, A.; Calin, M.; Zhang, L.C.; Prashanth, K.G.; Scudino, S.; Zhang, Y.S.; Eckert, J. Mechanical Behavior of Porous Commercially Pure Ti and Ti-TiB Composite Materials Manufactured by Selective Laser Melting. Mater. Sci. Eng. A 2015, 625, 350–356. [Google Scholar] [CrossRef]
  57. Zhuravleva, K.; Bönisch, M.; Prashanth, K.G.; Hempel, U.; Helth, A.; Gemming, T.; Calin, M.; Scudino, S.; Schultz, L.; Eckert, J.; et al. Production of Porous β-Type Ti-40Nb Alloy for Biomedical Applications: Comparison of Selective Laser Melting and Hot Pressing. Materials 2013, 6, 5700–5712. [Google Scholar] [CrossRef]
  58. Fisher, J.P.; Mikos, A.G.; Bronzino, J.D. Tissue Engineering. Science 1993, 260, 1–583. [Google Scholar] [CrossRef]
  59. Hutmacher, D.W. Scaffolds in Tissue Engineering Bone and Cartilage. Biomaterials 2000, 21, 2529–2543. [Google Scholar] [CrossRef]
  60. Pałka, K.; Pokrowiecki, R. Porous Titanium Implants: A Review. Adv. Eng. Mater. 2018, 20, 1700648. [Google Scholar] [CrossRef]
  61. Koju, N.; Niraula, S.; Fotovvati, B. Additively Manufactured Porous Ti6Al4V for Bone Implants: A Review. Metals 2022, 12, 687. [Google Scholar] [CrossRef]
  62. Niinomi, M.; Nakai, M.; Hieda, J. Development of New Metallic Alloys for Biomedical Applications. Acta Biomater. 2012, 8, 3888–3903. [Google Scholar] [CrossRef] [PubMed]
  63. Frost, H.M. Bone’s Mechanostat: A 2003 Update. Anat. Rec. Part A Discov. Mol. Cell. Evol. Biol. 2003, 275, 1081–1101. [Google Scholar] [CrossRef] [PubMed]
  64. Gardner, M.J.; Van Der Meulen, M.C.H.; Demetrakopoulos, D.; Wright, T.M.; Myers, E.R.; Bostrom, M.P. In Vivo Cyclic Axial Compression Affects Bone Healing in the Mouse Tibia. J. Orthop. Res. 2006, 24, 1679–1686. [Google Scholar] [CrossRef] [PubMed]
  65. Bozec, L.; Horton, M.A. Skeletal Tissues as Nanomaterials. J. Mater. Sci. Mater. Med. 2006, 17, 1043–1048. [Google Scholar] [CrossRef]
  66. Long, M.; Rack, H.J. Titanium Alloys in Total Joint Replacement—A Materials Science Perspective. Biomaterials 1998, 19, 1621–1639. [Google Scholar] [CrossRef]
  67. Sumner, D.R.; Turner, T.M.; Igloria, R.; Urban, R.M.; Galante, J.O. Functional Adaptation and Ingrowth of Bone Vary as a Function of Hip Implant Stiffness. J. Biomech. 1998, 31, 909–917. [Google Scholar] [CrossRef]
  68. Zhao, D.; Witte, F.; Lu, F.; Wang, J.; Li, J.; Qin, L. Current Status on Clinical Applications of Magnesium-Based Orthopaedic Implants: A Review from Clinical Translational Perspective. Biomaterials 2017, 112, 287–302. [Google Scholar] [CrossRef]
  69. Agarwal, S.; Curtin, J.; Duffy, B.; Jaiswal, S. Biodegradable Magnesium Alloys for Orthopaedic Applications: A Review on Corrosion, Biocompatibility and Surface Modifications. Mater. Sci. Eng. C 2016, 68, 948–963. [Google Scholar] [CrossRef]
  70. Osuchukwu, O.A.; Salihi, A.; Abdullahi, I.; Obada, D.O.; Abolade, S.A.; Akande, A.; Csaki, S.; Dodoo-Arhin, D. Datasets on the Elastic and Mechanical Properties of Hydroxyapatite: A First Principle Investigation, Experiments, and Pedagogical Perspective. Data Brief 2023, 48, 109075. [Google Scholar] [CrossRef]
  71. Piconi, C.; Maccauro, G. Zirconia as a Ceramic Biomaterial. Biomaterials 1999, 20, 1–25. [Google Scholar] [CrossRef]
  72. Attar, H.; Prashanth, K.G.; Chaubey, A.K.; Calin, M.; Zhang, L.C.; Scudino, S.; Eckert, J. Comparison of Wear Properties of Commercially Pure Titanium Prepared by Selective Laser Melting and Casting Processes. Mater. Lett. 2015, 142, 38–41. [Google Scholar] [CrossRef]
  73. Yan, Q.; Hou, L.; Sun, J.; Li, D.; Song, J.; Liu, X.; Wang, Q.; Wei, Y. Regulating the Localized Corrosion of Grain Boundary and Galvanic Corrosion by Adding the Electronegative Element in Magnesium Alloy. Corros. Sci. 2025, 244, 112667. [Google Scholar] [CrossRef]
  74. Zhang, T.; Wang, Z.; Qiu, Y.; Iftikhar, T.; Liu, H. “Electrons-Siphoning” of Sulfate Reducing Bacteria Biofilm Induced Sharp Depletion of Al-Zn-In-Mg-Si Sacrificial Anode in the Galvanic Corrosion Coupled with Carbon Steel. Corros. Sci. 2023, 216, 111103. [Google Scholar] [CrossRef]
  75. Nomura, K.; Sakairi, M.; Fushimi, K. Ion-Selectivity of Galvanic Corrosion Products Formed in Multi-Material Gaps under Atmospheric Corrosion Conditions with Anti-Freezing Salts. Corros. Sci. 2024, 237, 112305. [Google Scholar] [CrossRef]
  76. Lucas, L.C.; Buchanan, R.A.; Lemons, J.E. Investigations on the Galvanic Corrosion of Multialloy Total Hip Prostheses. J. Biomed. Mater. Res. 1981, 15, 731–747. [Google Scholar] [CrossRef] [PubMed]
  77. Szklarska-Smialowska, Z. Pitting Corrosion of Aluminum. Corros. Sci. 1999, 41, 1743–1767. [Google Scholar] [CrossRef]
  78. Loto, R.T. Pitting Corrosion Resistance and Inhibition of Lean Austenitic Stainless Steel Alloys. Austenitic Stainl. Steels New Asp. 2017, 147–170. [Google Scholar] [CrossRef]
  79. Zhao, K.; Liu, J.H.; Yu, M.; Li, S.M. Through-Thickness Inhomogeneity of Precipitate Distribution and Pitting Corrosion Behavior of Al–Li Alloy Thick Plate. Trans. Nonferrous Met. Soc. China 2019, 29, 1793–1802. [Google Scholar] [CrossRef]
  80. McCafferty, E. Introduction to Corrosion Science; Springer: Berlin/Heidelberg, Germany, 2010; ISBN 9781441904546. [Google Scholar]
  81. Prashanth, K.G.; Debalina, B.; Wang, Z.; Gostin, P.F.; Gebert, A.; Calin, M.; Kühn, U.; Kamaraj, M.; Scudino, S.; Eckert, J. Tribological and Corrosion Properties of Al-12Si Produced by Selective Laser Melting. J. Mater. Res. 2014, 29, 2044–2054. [Google Scholar] [CrossRef]
  82. Aoyama, T.; Kato, C. Introduction of Cu2+ to the inside of the Crevice by Chelation and Its Effect on Crevice Corrosion of Type 316L Stainless Steel. Corros. Sci. 2023, 210, 110850. [Google Scholar] [CrossRef]
  83. Zhao, M.; Wang, J.; Ma, Y.; Xu, Y.; Zhang, X.; Li, Z.; Xu, D.; Wang, F. Electroactive Shewanella Algae Accelerates the Crevice Corrosion of X70 Pipeline Steel in Marine Environment. Corros. Sci. 2024, 235, 112226. [Google Scholar] [CrossRef]
  84. Liu, H.; Wang, Z.; Qiao, L.; Su, Y.; Yan, Y. Role of Protein in Crevice Corrosion of CoCrMo Alloy: An Investigation Using Wire Beam Electrodes. Corros. Sci. 2023, 215, 111028. [Google Scholar] [CrossRef]
  85. Levine, D.L.; Staehle, R.W. Crevice Corrosion in Orthopedic Implant Metals. J. Biomed. Mater. Res. 1977, 11, 553–561. [Google Scholar] [CrossRef] [PubMed]
  86. Uva Narayanan, C.; Daniel, A.; Praveenkumar, K.; Manivasagam, G.; Suwas, S.; Prashanth, K.G.; Suya Prem Anand, P. Effect of Scanning Speed on Mechanical, Corrosion, and Fretting-Tribocorrosion Behavior of Austenitic 316L Stainless Steel Produced by Laser Powder Bed Fusion Process. J. Manuf. Process 2024, 131, 1582–1593. [Google Scholar] [CrossRef]
  87. Kang, L.Z.; Lu, Y.H.; Bian, W.W.; Yu, P.J.; Wang, Y.B.; Xin, L.; Han, Y.M. Fretting Corrosion Behavior and Microstructure Evolution of Hydrided Zirconium Alloy under Gross Slip Regime in High Temperature High Pressure Water Environment. Corros. Sci. 2025, 242, 112582. [Google Scholar] [CrossRef]
  88. Bian, W.W.; Lu, Y.H.; Zhang, X.F.; Han, Y.M.; Wang, F.; Shoji, T. Effect of Fretting Wear Regimes on Stress Corrosion Cracking of Alloy 690TT in High-Temperature Pressurized Water. Corros. Sci. 2024, 237, 112320. [Google Scholar] [CrossRef]
  89. Sargeant, A.; Goswami, T. Hip Implants: Paper V. Physiological Effects. Mater. Des. 2006, 27, 287–307. [Google Scholar] [CrossRef]
  90. Ghosh, S.; Choudhury, D.; Roy, T.; Moradi, A.; Masjuki, H.H.; Pingguan-Murphy, B. Tribological Performance of the Biological Components of Synovial Fluid in Artificial Joint Implants. Sci. Technol. Adv. Mater. 2015, 16, 045002. [Google Scholar] [CrossRef]
  91. Eliaz, N. Corrosion of Metallic Biomaterials: A Review. Materials 2019, 12, 407. [Google Scholar] [CrossRef]
  92. Seo, E.; Lee, Y.N.; Shin, W.Y.; Kim, K.H.; Jung, S.H.; Kang, H.G.; Kim, R.; Sung, H.; Jung, I.D.; Park, J.W. Structural Influence on Titanium Ion Dissolution in 3D-Printed Ti6Al4V Orthopedic Implants. Sci. Rep. 2025, 15, 1–18. [Google Scholar] [CrossRef]
  93. Hermawan, H.; Razavi, M. Absorbable Metals for Biomedical Applications; MDPI-Multidisciplinary Digital Publishing Institute: Basel, Switzerland, 2021; ISBN 9783036517643. [Google Scholar]
  94. Nair, M.M.; Swaroop, S. Impact of the Oxide Layer and Subsurface Micromechanical Properties of Laser Peened 31L Stainless Steel on Biocorrosion Resistance in Simulated Body Fluid. Surf. Interfaces 2025, 56, 105672. [Google Scholar] [CrossRef]
  95. Prashanth, K.G. Interpenetrating Composites: A Nomenclature Dilemma. Materials 2025, 18, 273. [Google Scholar] [CrossRef] [PubMed]
  96. Zhang, Z.; Wang, Z.; Zhao, Q.; Prashanth, K.G. Metal-Metal Interpenetrating Phase Composites: A Review. J. Alloys Compd. 2024, 1009, 176951. [Google Scholar] [CrossRef]
  97. Abbasi, N.; Hamlet, S.; Love, R.M.; Nguyen, N.T. Porous Scaffolds for Bone Regeneration. J. Sci. Adv. Mater. Devices 2020, 5, 1–9. [Google Scholar] [CrossRef]
  98. Kuboki, Y.; Takita, H.; Kobayashi, D.; Tsuruga, E.; Inoue, M.; Murata, M.; Nagai, N.; Dohi, Y.; Ohgushi, H. BMP-Induced Osteogenesis on the Surface of Hydroxyapatite with Geometrically Feasible and Nonfeasible Structures: Topology of Osteogenesis. J. Biomed. Mater. Res. 1998, 39, 190–199. [Google Scholar] [CrossRef]
  99. Bai, F.; Wang, Z.; Lu, J.; Liu, J.; Chen, G.; Lv, R.; Wang, J.; Lin, K.; Zhang, J.; Huang, X. The Correlation between the Internal Structure and Vascularization of Controllable Porous Bioceramic Materials in Vivo: A Quantitative Study. Tissue Eng. Part A 2010, 16, 3791–3803. [Google Scholar] [CrossRef]
  100. Lu, J.X.; Flautre, B.; Anselme, K.; Hardouin, P.; Gallur, A.; Descamps, M.; Thierry, B. Role of Interconnections in Porous Bioceramics on Bone Recolonization in Vitro and in Vivo. J. Mater. Sci. Mater. Med. 1999, 10, 111–120. [Google Scholar] [CrossRef]
  101. Abdel-Hady Gepreel, M.; Niinomi, M. Biocompatibility of Ti-Alloys for Long-Term Implantation. J. Mech. Behav. Biomed. Mater. 2013, 20, 407–415. [Google Scholar] [CrossRef]
  102. Sauer, B.W.; Weinstein, A.M.; Klawitter, J.J.; Hulbert, S.F.; Leonard, R.B.; Bagwell, J.G. The Role of Porous Polymeric Materials in Prosthesis Attachment. J. Biomed. Mater. Res. 1974, 8, 145–153. [Google Scholar] [CrossRef]
  103. 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]
  104. Laptev, A.; Vyal, O.; Bram, M.; Buchkremer, H.P.; Stöver, D. Green Strength of Powder Compacts Provided Production of Highly Porous Titanium Parts. Powder Metall. 2005, 48, 358–364. [Google Scholar] [CrossRef]
  105. Gittens, R.A.; Scheideler, L.; Rupp, F.; Hyzy, S.L.; Geis-Gerstorfer, J.; Schwartz, Z.; Boyan, B.D. A Review on the Wettability of Dental Implant Surfaces II: Biological and Clinical Aspects. Acta Biomater. 2014, 10, 2907–2918. [Google Scholar] [CrossRef] [PubMed]
  106. Peta, K.; Kubiak, K.J.; Sfravara, F.; Brown, C.A. Dynamic Wettability of Complex Fractal Isotropic Surfaces—Multiscale Correlations. Tribol. Int. 2026, 214, 111145. [Google Scholar] [CrossRef]
  107. Peta, K.; Bartkowiak, T.; Rybicki, M.; Galek, P.; Mendak, M.; Wieczorowski, M.; Brown, C.A. Scale-Dependent Wetting Behavior of Bioinspired Lubricants on Electrical Discharge Machined Ti6Al4V Surfaces. Tribol. Int. 2024, 194, 109562. [Google Scholar] [CrossRef]
  108. Gittens, R.A.; Olivares-Navarrete, R.; Cheng, A.; Anderson, D.M.; McLachlan, T.; Stephan, I.; Geis-Gerstorfer, J.; Sandhage, K.H.; Fedorov, A.G.; Rupp, F.; et al. The Roles of Titanium Surface Micro/Nanotopography and Wettability on the Differential Response of Human Osteoblast Lineage Cells. Acta Biomater. 2013, 9, 6268–6277. [Google Scholar] [CrossRef]
  109. Gittens, R.A.; Olivares-Navarrete, R.; McLachlan, T.; Cai, Y.; Hyzy, S.L.; Schneider, J.M.; Schwartz, Z.; Sandhage, K.H.; Boyan, B.D. Differential Responses of Osteoblast Lineage Cells to Nanotopographically-Modified, Microroughened Titanium–Aluminum–Vanadium Alloy Surfaces. Biomaterials 2012, 33, 8986–8994. [Google Scholar] [CrossRef]
  110. Gittens, R.A.; McLachlan, T.; Olivares-Navarrete, R.; Cai, Y.; Berner, S.; Tannenbaum, R.; Schwartz, Z.; Sandhage, K.H.; Boyan, B.D. The Effects of Combined Micron-/Submicron-Scale Surface Roughness and Nanoscale Features on Cell Proliferation and Differentiation. Biomaterials 2011, 32, 3395–3403. [Google Scholar] [CrossRef]
  111. Huang, Q.; Lin, L.; Yang, Y.; Hu, R.; Vogler, E.A.; Lin, C. Role of Trapped Air in the Formation of Cell-and-Protein Micropatterns on Superhydrophobic/Superhydrophilic Microtemplated Surfaces. Biomaterials 2012, 33, 8213–8220. [Google Scholar] [CrossRef]
  112. Park, J.H.; Schwartz, Z.; Olivares-Navarrete, R.; Boyan, B.D.; Tannenbaum, R. Enhancement of Surface Wettability via the Modification of Microtextured Titanium Implant Surfaces with Polyelectrolytes. Langmuir 2011, 27, 5976–5985. [Google Scholar] [CrossRef]
  113. Viceconti, M.; Muccini, R.; Bernakiewicz, M.; Baleani, M.; Cristofolini, L. Large-sliding contact elements accurately predict levels of bone–implant micromotion relevant to osseointegration. J. Biomech. 2000, 33, 1611–1618. [Google Scholar] [CrossRef]
  114. Van Noort, R. Titanium: The Implant Material of Today. J. Mater. Sci. 1987, 22, 3801–3811. [Google Scholar] [CrossRef]
  115. Hallab, N.J.; Anderson, S.; Stafford, T.; Glant, T.; Jacobs, J.J. Lymphocyte Responses in Patients with Total Hip Arthroplasty. J. Orthop. Res. 2005, 23, 384391. [Google Scholar] [CrossRef] [PubMed]
  116. Donaruma, L.G. Definitions in Biomaterials, D.F. Williams, Ed., Elsevier, Amsterdam, 1987, 72 pp. J. Polym. Sci. Part C Polym. Lett. 1988, 26, 414. [Google Scholar] [CrossRef]
  117. Ratner, B.D. The Biocompatibility of Implant Materials; Elsevier: Amsterdam, The Netherlands, 2015; ISBN 9780128005002. [Google Scholar]
  118. Williams, D.F. On the Mechanisms of Biocompatibility. Biomaterials 2008, 29, 2941–2953. [Google Scholar] [CrossRef]
  119. Manivasagam, G.; Dhinasekaran, D.; Rajamanickam, A. Biomedical Implants: Corrosion and Its Prevention—A Review. Recent Pat. Corros. Science 2010, 2, 40–54. [Google Scholar] [CrossRef]
  120. Zidrou, C.; Kapetanou, A.; Rizou, S. The Effect of Drugs on Implant Osseointegration—A Narrative Review. Injury 2023, 54, 110888. [Google Scholar] [CrossRef]
  121. Jin, W.; Chu, P.K. Orthopedic Implants. Encycl. Biomed. Eng. 2019, 1, 425–439. [Google Scholar] [CrossRef]
  122. Mello-Machado, R.C.; Sartoretto, S.C.; Granjeiro, J.M.; Calasans-Maia, J.d.A.; de Uzeda, M.J.P.G.; Mourão, C.F.d.A.B.; Ghiraldini, B.; Bezerra, F.J.B.; Senna, P.M.; Calasans-Maia, M.D. Osseodensification Enables Bone Healing Chambers with Improved Low-Density Bone Site Primary Stability: An in Vivo Study. Sci. Rep. 2021, 11, 15436. [Google Scholar] [CrossRef]
  123. Martin, R.B.; Burr, D.B. Structure, Function and Adaptation of Compact Bone. Skelet. Radiol. 1989, 18, 506. [Google Scholar] [CrossRef]
  124. Siddiqui, S.A.; Srikanth, S.P.; Wu, Y.S.; Kalita, T.; Ambartsumov, T.G.; Tseng, W.; Kumar, A.P.; Ahmad, A.; Michalek, J.E. Different Types of Algae Beneficial for Bone Health in Animals and in Humans—A Review. Algal. Res. 2024, 82, 103593. [Google Scholar] [CrossRef]
  125. Lakes, R.S.; Park, J. Biomaterials. An Introduction, 2nd ed.; Springer: Berlin/Heidelberg, Germany, 1993; ISBN 0-306-43992-1. [Google Scholar]
  126. Fung, Y.-C. Biomechanics; Springer: Berlin/Heidelberg, Germany, 1993. [Google Scholar] [CrossRef]
  127. Lopes, D.; Martins-Cruz, C.; Oliveira, M.B.; Mano, J.F. Bone Physiology as Inspiration for Tissue Regenerative Therapies. Biomaterials 2018, 185, 240–275. [Google Scholar] [CrossRef] [PubMed]
  128. Kaplan, F.S.; Hayes, W.C.; Keaveny, T.M.; Boskey, A.; Einhorn, T.A.; Iannotti, J.P. Form and Function of Bone. Orthop. Basic Sci. 1994, 4, 127–185. [Google Scholar]
  129. Boskey, A.L.; Coleman, R. Aging and Bone. J. Dent. Res. 2015, 89, 1333–1348. [Google Scholar] [CrossRef] [PubMed]
  130. Raja, I.S.; Preeth, D.R.; Vedhanayagam, M.; Hyon, S.H.; Lim, D.; Kim, B.; Rajalakshmi, S.; Han, D.W. Polyphenols-Loaded Electrospun Nanofibers in Bone Tissue Engineering and Regeneration. Biomater. Res. 2021, 25, 29. [Google Scholar] [CrossRef]
  131. Webster, T.J. Nanophase Ceramics: The Future Orthopedic and Dental Implant Material. Adv. Chem. Eng. 2001, 27, 125–166. [Google Scholar] [CrossRef]
  132. Friedenstein, A.J.; Chailakhyan, R.K.; Gerasimov, U.V. Bone Marrow Osteogenic Stem Cells: In Vitro Cultivation and Transplantation in Diffusion Chambers. Cell Prolif. 1987, 20, 263–272. [Google Scholar] [CrossRef]
  133. Liu, J.; Liu, D.; Li, S.; Deng, Z.; Pan, Z.; Li, C.; Chen, T. The Effects of Graphene Oxide Doping on the Friction and Wear Properties of TiN Bioinert Ceramic Coatings Prepared Using Wide-Band Laser Cladding. Surf. Coat. Technol. 2023, 458, 129354. [Google Scholar] [CrossRef]
  134. Desante, G.; Labude, N.; Rütten, S.; Römer, S.; Kaufmann, R.; Zybała, R.; Jagiełło, J.; Lipińska, L.; Chlanda, A.; Telle, R.; et al. Graphene Oxide Nanofilm to Functionalize Bioinert High Strength Ceramics. Appl. Surf. Sci. 2021, 566, 150670. [Google Scholar] [CrossRef]
  135. Chen, T.; Deng, Z.; Liu, D.; Zhu, X.; Xiong, Y. Bioinert TiC Ceramic Coating Prepared by Laser Cladding: Microstructures, Wear Resistance, and Cytocompatibility of the Coating. Surf. Coat. Technol. 2021, 423, 127635. [Google Scholar] [CrossRef]
  136. Bohner, M. Bioresorbable Ceramics. In Degradation Rate of Bioresorbable Materials: Prediction and Evaluation; CRC Press: Boca Raton, FL, USA, 2008; pp. 95–114. [Google Scholar] [CrossRef]
  137. Tian, X.; Raina, D.B.; Oberländer, J.T.; Liu, Y.; Goronzy, J.; Apolle, R.; Vater, C.; Richter, R.F.; Tägil, M.; Lidgren, L.; et al. Comparison of Immediate Anchoring Effectiveness of Two Different Techniques of Bioresorbable Ceramic Application for Pedicle Screw Augmentation. Ceram. Int. 2024, 50, 12877–12889. [Google Scholar] [CrossRef]
  138. Wang, X.; Aziz, A.; Sheng, X.; Wang, L.; Yin, L. Bioresorbable Neural Interfaces for Bioelectronic Medicine. Curr. Opin. Biomed. Eng. 2024, 32, 100565. [Google Scholar] [CrossRef]
  139. Mehboob, H.; Chang, S.H. Application of Composites to Orthopedic Prostheses for Effective Bone Healing: A Review. Compos. Struct. 2014, 118, 328–341. [Google Scholar] [CrossRef]
  140. Ibrahim, A.M.S.; Koolen, P.G.L.; Kim, K.; Perrone, G.S.; Kaplan, D.L.; Lin, S.J. Absorbable Biologically Based Internal Fixation. Clin. Podiatr. Med. Surg. 2015, 32, 61–72. [Google Scholar] [CrossRef] [PubMed]
  141. Lee, H.; Zeng, F.; Dunne, M.; Allen, C. Methoxy Poly(Ethylene Glycol)-Block-Poly(δ-Valerolactone) Copolymer Micelles for Formulation of Hydrophobic Drugs. Biomacromolecules 2005, 6, 3119–3128. [Google Scholar] [CrossRef]
  142. Kim, S.C.; Kim, D.W.; Shim, Y.H.; Bang, J.S.; Oh, H.S.; Kim, S.W.; Seo, M.H. In Vivo Evaluation of Polymeric Micellar Paclitaxel Formulation: Toxicity and Efficacy. J. Control. Release 2001, 72, 191–202. [Google Scholar] [CrossRef]
  143. Lausmaa, J.; Kasemo, B.; Mattsson, H.; Odelius, H. Multi-Technique Surface Charaterization of Oxide Films on Electropolished and Anodically Oxidized Titanium. Appl. Surf. Sci. 1990, 45, 189–200. [Google Scholar] [CrossRef]
  144. Gu, X.; Zheng, Y.; Cheng, Y.; Zhong, S.; Xi, T. In Vitro Corrosion and Biocompatibility of Binary Magnesium Alloys. Biomaterials 2009, 30, 484–498. [Google Scholar] [CrossRef]
  145. Chen, Y.J.; Li, Y.J.; Walmsley, J.C.; Dumoulin, S.; Skaret, P.C.; Roven, H.J. Microstructure Evolution of Commercial Pure Titanium during Equal Channel Angular Pressing. Mater. Sci. Eng. A 2010, 527, 789–796. [Google Scholar] [CrossRef]
  146. Barceloux, D.G.; Barceloux, D. Chromium. J. Toxicol. Clin. Toxicol. 1999, 37, 173–194. [Google Scholar] [CrossRef]
  147. Barceloux, D.G.; Barceloux, D. Nickel. J. Toxicol. Clin. Toxicol. 1999, 37, 239–258. [Google Scholar] [CrossRef]
  148. Li, X.; Gao, P.; Wan, P.; Pei, Y.; Shi, L.; Fan, B.; Shen, C.; Xiao, X.; Yang, K.; Guo, Z. Novel Bio-Functional Magnesium Coating on Porous Ti6Al4V Orthopaedic Implants: In Vitro and In Vivo Study. Sci. Rep. 2017, 7, 1–11. [Google Scholar] [CrossRef] [PubMed]
  149. He, M.; Chen, L.; Yin, M.; Xu, S.; Liang, Z. Review on Magnesium and Magnesium-Based Alloys as Biomaterials for Bone Immobilization. J. Mater. Res. Technol. 2023, 23, 4396–4419. [Google Scholar] [CrossRef]
  150. Badkoobeh, F.; Mostaan, H.; Rafiei, M.; Bakhsheshi-Rad, H.R.; RamaKrishna, S.; Chen, X. Additive Manufacturing of Biodegradable Magnesium-Based Materials: Design Strategies, Properties, and Biomedical Applications. J. Magnes. Alloys 2023, 11, 801–839. [Google Scholar] [CrossRef]
  151. Yadav, M.; Shukla, R.; Kesavan, P.; Nilawar, S.; Perugu, C.; Sellamuthu, P.; Chatterjee, K.; Suwas, S.; Jayamani, J.; Prashanth, K. Microstructural, Mechanical, Corrosion, and Biological Behavior of Spark Plasma Sintered Commercially Pure Zinc for Biomedical Applications. Mater. Adv. 2025, 6, 3546–3560. [Google Scholar] [CrossRef]
  152. Eddy Jai Poinern, G.; Brundavanam, S.; Fawcett, D. Biomedical Magnesium Alloys: A Review of Material Properties, Surface Modifications and Potential as a Biodegradable Orthopaedic Implant. Am. J. Biomed. Eng. 2013, 2, 218–240. [Google Scholar] [CrossRef]
  153. Murugan, R.; Ramakrishna, S. Development of Nanocomposites for Bone Grafting. Compos. Sci. Technol. 2005, 65, 2385–2406. [Google Scholar] [CrossRef]
  154. Zhao, Y.; Wu, G.; Jiang, J.; Wong, H.M.; Yeung, K.W.K.; Chu, P.K. Improved Corrosion Resistance and Cytocompatibility of Magnesium Alloy by Two-Stage Cooling in Thermal Treatment. Corros. Sci. 2012, 59, 360–365. [Google Scholar] [CrossRef]
  155. Witte, F.; Kaese, V.; Haferkamp, H.; Switzer, E.; Meyer-Lindenberg, A.; Wirth, C.J.; Windhagen, H. In Vivo Corrosion of Four Magnesium Alloys and the Associated Bone Response. Biomaterials 2005, 26, 3557–3563. [Google Scholar] [CrossRef]
  156. Ghali, E. Corrosion Resistance of Aluminum and Magnesium Alloys: Understanding, Performance, and Testing. In Corrosion Resistance of Aluminum and Magnesium Alloys: Understanding, Performance, and Testing; John Wiley & Sons: Hoboken, NJ, USA, 2010. [Google Scholar] [CrossRef]
  157. Radha, R.; Sreekanth, D. Insight of Magnesium Alloys and Composites for Orthopedic Implant Applications—A Review. J. Magnes. Alloys 2017, 5, 286–312. [Google Scholar] [CrossRef]
  158. Leyens, C.; Peters, M. Titanium and Titanium Alloys: Fundamentals and Applications; Wiley-VCH: Weinheim, Germany, 2003. [Google Scholar] [CrossRef]
  159. Banerjee, D.; Williams, J.C. Perspectives on Titanium Science and Technology. Acta Mater. 2013, 61, 844–879. [Google Scholar] [CrossRef]
  160. Singh, R.; Lee, P.D.; Dashwood, R.J.; Lindley, T.C. Titanium Foams for Biomedical Applications: A Review. Mater. Technol. 2010, 25, 127–136. [Google Scholar] [CrossRef]
  161. Qian, M.; Xu, W.; Brandt, M.; Tang, H.P. Additive Manufacturing and Postprocessing of Ti-6Al-4V for Superior Mechanical Properties. MRS Bull. 2016, 41, 775–783. [Google Scholar] [CrossRef]
  162. Patnaik, P. Handbook of Inorganic Chemicals; McGraw-Hill: New York, NY, USA, 2003; ISBN 0070494398. [Google Scholar]
  163. Kim, S.W.; Jung, H.D.; Kang, M.H.; Kim, H.E.; Koh, Y.H.; Estrin, Y. Fabrication of Porous Titanium Scaffold with Controlled Porous Structure and Net-Shape Using Magnesium as Spacer. Mater. Sci. Eng. C 2013, 33, 2808–2815. [Google Scholar] [CrossRef]
  164. Esen, Z.; Bor, Ş. Processing of Titanium Foams Using Magnesium Spacer Particles. Scr. Mater. 2007, 56, 341–344. [Google Scholar] [CrossRef]
  165. Wang, X.; Lu, Z.; Jia, L.; Li, F. Preparation and Properties of Low Cost Porous Titanium by Using Rice Husk as Hold Space. Prog. Nat. Sci. Mater. Int. 2017, 27, 344–349. [Google Scholar] [CrossRef]
  166. Wang, X.S.; Lu, Z.L.; Jia, L.; Chen, J.X. Preparation of Porous Titanium Materials by Powder Sintering Process and Use of Space Holder Technique. J. Iron Steel Res. Int. 2017, 24, 97–102. [Google Scholar] [CrossRef]
  167. Jian, X.; Yongning, L.; Yong, L.; Guibao, Q.; Jinming, L. The Application of Model Equation Method in Preparation of Titanium Foams. J. Mater. Res. Technol. 2021, 13, 121–127. [Google Scholar] [CrossRef]
  168. Vasconcellos, L.M.R.D.; Oliveira, M.V.D.; Graça, M.L.D.A.; Vasconcellos, L.G.O.D.; Carvalho, Y.R.; Cairo, C.A.A. Porous Titanium Scaffolds Produced by Powder Metallurgy for Biomedical Applications. Mater. Res. 2008, 11, 275–280. [Google Scholar] [CrossRef]
  169. Niu, W.; Bai, C.; Qiu, G.B.; Wang, Q. Processing and Properties of Porous Titanium Using Space Holder Technique. Mater. Sci. Eng. A 2009, 506, 148–151. [Google Scholar] [CrossRef]
  170. Li, D.S.; Zhang, Y.P.; Ma, X.; Zhang, X.P. Space-Holder Engineered Porous NiTi Shape Memory Alloys with Improved Pore Characteristics and Mechanical Properties. J. Alloys Compd. 2009, 474, L1. [Google Scholar] [CrossRef]
  171. Xiang, C.; Zhang, Y.; Li, Z.; Zhang, H.; Huang, Y.; Tang, H. Preparation and Compressive Behavior of Porous Titanium Prepared by Space Holder Sintering Process. Procedia Eng. 2012, 27, 768–774. [Google Scholar] [CrossRef]
  172. Torres, Y.; Pavón, J.J.; Rodríguez, J.A. Processing and Characterization of Porous Titanium for Implants by Using NaCl as Space Holder. J. Mater. Process. Technol. 2012, 212, 1061–1069. [Google Scholar] [CrossRef]
  173. Chen, Y.; Kent, D.; Bermingham, M.; Dehghan-Manshadi, A.; Dargusch, M. Manufacturing of Biocompatible Porous Titanium Scaffolds Using a Novel Spherical Sugar Pellet Space Holder. Mater. Lett. 2017, 195, 92–95. [Google Scholar] [CrossRef]
  174. Li, B.Q.; Wang, C.Y.; Lu, X. Effect of Pore Structure on the Compressive Property of Porous Ti Produced by Powder Metallurgy Technique. Mater. Des. 2013, 50, 613–619. [Google Scholar] [CrossRef]
  175. Wang, D.; Li, Q.; Xu, M.; Jiang, G.; Zhang, Y.; He, G. A Novel Approach to Fabrication of Three-Dimensional Porous Titanium with Controllable Structure. Mater. Sci. Eng. C 2017, 71, 1046–1051. [Google Scholar] [CrossRef]
  176. Chen, Y.; Frith, J.E.; Dehghan-Manshadi, A.; Attar, H.; Kent, D.; Soro, N.D.M.; Bermingham, M.J.; Dargusch, M.S. Mechanical Properties and Biocompatibility of Porous Titanium Scaffolds for Bone Tissue Engineering. J. Mech. Behav. Biomed. Mater. 2017, 75, 169–174. [Google Scholar] [CrossRef]
  177. Zou, C.; Zhang, E.; Li, M.; Zeng, S. Preparation, Microstructure and Mechanical Properties of Porous Titanium Sintered by Ti Fibres. J. Mater. Sci. Mater. Med. 2008, 19, 401–405. [Google Scholar] [CrossRef]
  178. Wolf, F.I.; Trapani, V. Cell (Patho)Physiology of Magnesium. Clin. Sci. 2008, 114, 27–35. [Google Scholar] [CrossRef]
  179. Maguire, M.E.; Cowan, J.A. Magnesium Chemistry and Biochemistry. BioMetals 2002, 15, 203–210. [Google Scholar] [CrossRef]
  180. Vormann, J. Magnesium: Nutrition and Metabolism. Mol. Aspects Med. 2003, 24, 27–37. [Google Scholar] [CrossRef]
  181. Flatman, P.W. Magnesium Transport across Cell Membranes. J. Membr. Biol. 1984, 80, 1–14. [Google Scholar] [CrossRef] [PubMed]
  182. Cachinho, S.C.P.; Correia, R.N. Titanium Scaffolds for Osteointegration: Mechanical, in Vitro and Corrosion Behaviour. J. Mater. Sci. Mater. Med. 2008, 19, 451–457. [Google Scholar] [CrossRef]
  183. Karageorgiou, V.; Kaplan, D. Porosity of 3D Biomaterial Scaffolds and Osteogenesis. Biomaterials 2005, 26, 5474–5491. [Google Scholar] [CrossRef] [PubMed]
  184. Barrabés, M.; Sevilla, P.; Planell, J.A.; Gil, F.J. Mechanical Properties of Nickel–Titanium Foams for Reconstructive Orthopaedics. Mater. Sci. Eng. C 2008, 28, 23–27. [Google Scholar] [CrossRef]
  185. Li, J.P.; Li, S.H.; De Groot, K.; Layrolle, P. Preparation and Characterization of Porous Titanium. Key Eng. Mater. 2002, 218, 51–54. [Google Scholar] [CrossRef]
  186. Wang, C.; Chen, H.; Zhu, X.; Xiao, Z.; Zhang, K.; Zhang, X. An Improved Polymeric Sponge Replication Method for Biomedical Porous Titanium Scaffolds. Mater. Sci. Eng. C 2017, 70, 1192–1199. [Google Scholar] [CrossRef]
  187. He, G.; Liu, P.; Tan, Q. Porous Titanium Materials with Entangled Wire Structure for Load-Bearing Biomedical Applications. J. Mech. Behav. Biomed. Mater. 2012, 5, 16–31. [Google Scholar] [CrossRef]
  188. Jiang, G.; He, G. Enhancement of the Porous Titanium with Entangled Wire Structure for Load-Bearing Biomedical Applications. Mater. Des. 2014, 56, 241–244. [Google Scholar] [CrossRef]
  189. Tan, Q.; Liu, P.; Du, C.; Wu, L.; He, G. Mechanical Behaviors of Quasi-Ordered Entangled Aluminum Alloy Wire Material. Mater. Sci. Eng. A 2009, 527, 38–44. [Google Scholar] [CrossRef]
  190. Liu, P.; Tan, Q.; Wu, L.; He, G. Compressive and Pseudo-Elastic Hysteresis Behavior of Entangled Titanium Wire Materials. Mater. Sci. Eng. A 2010, 527, 3301–3309. [Google Scholar] [CrossRef]
  191. Jiang, G.; Wang, C.; Li, Q.; Dong, J.; He, G. Porous Titanium with Entangled Structure Filled with Biodegradable Magnesium for Potential Biomedical Applications. Mater. Sci. Eng. C 2015, 47, 142–149. [Google Scholar] [CrossRef] [PubMed]
  192. Liu, Y.; Jiang, G.; He, G. Enhancement of Entangled Porous Titanium by BisGMA for Load-Bearing Biomedical Applications. Mater. Sci. Eng. C 2016, 61, 37–41. [Google Scholar] [CrossRef] [PubMed]
  193. He, G.; Liu, P.; Tan, Q.; Jiang, G. Flexural and Compressive Mechanical Behaviors of the Porous Titanium Materials with Entangled Wire Structure at Different Sintering Conditions for Load-Bearing Biomedical Applications. J. Mech. Behav. Biomed. Mater. 2013, 28, 309–319. [Google Scholar] [CrossRef] [PubMed]
  194. Asaoka, K.; Kuwayama, N.; Okuno, O.; Miura, I. Mechanical Properties and Biomechanical Compatibility of Porous Titanium for Dental Implants. J. Biomed. Mater. Res. 1985, 19, 699–713. [Google Scholar] [CrossRef]
  195. Ivasishin, O.M.; Savvakin, D.G. The Impact of Diffusion on Synthesis of High-Strength Titanium Alloys from Elemental Powder Blends. Key Eng. Mater. 2010, 436, 113–121. [Google Scholar] [CrossRef]
  196. Yang, Y.F.; Luo, S.D.; Schaffer, G.B.; Qian, M. Sintering of Ti–10V–2Fe–3Al and Mechanical Properties. Mater. Sci. Eng. A 2011, 528, 6719–6726. [Google Scholar] [CrossRef]
  197. Mamedov, V. Spark Plasma Sintering as Advanced PM Sintering Method. Powder Metall. 2002, 45, 322–328. [Google Scholar] [CrossRef]
  198. Xie, G.; Ohashi, O.; Chiba, K.; Yamaguchi, N.; Song, M.; Furuya, K.; Noda, T. Frequency Effect on Pulse Electric Current Sintering Process of Pure Aluminum Powder. Mater. Sci. Eng. A 2003, 359, 384–390. [Google Scholar] [CrossRef]
  199. Xie, G.; Ohashi, O.; Wada, K.; Ogawa, T.; Song, M.; Furuya, K. Interface Microstructure of Aluminum Die-Casting Alloy Joints Bonded by Pulse Electric-Current Bonding Process. Mater. Sci. Eng. A 2006, 428, 12–17. [Google Scholar] [CrossRef]
  200. Wang, S.W.; Chen, L.D.; Kang, Y.S.; Niino, M.; Hirai, T. Effect of Plasma Activated Sintering (PAS) Parameters on Densification of Copper Powder. Mater. Res. Bull. 2000, 35, 619–628. [Google Scholar] [CrossRef]
  201. Yang, Y.F.; Qian, M. Spark Plasma Sintering and Hot Pressing of Titanium and Titanium Alloys. In Titanium Powder Metallurgy: Science, Technology and Applications; Butterworth-Heinemann: Oxford, UK, 2015; pp. 219–235. [Google Scholar] [CrossRef]
  202. Omori, M. Sintering, Consolidation, Reaction and Crystal Growth by the Spark Plasma System (SPS). Mater. Sci. Eng. A 2000, 287, 183–188. [Google Scholar] [CrossRef]
  203. Groza, J.R.; Zavaliangos, A. Sintering Activation by External Electrical Field. Mater. Sci. Eng. A 2000, 287, 171–177. [Google Scholar] [CrossRef]
  204. Eriksson, M.; Shen, Z.; Nygren, M. Fast Densification and Deformation of Titanium Powder. Powder Metall. 2013, 48, 231–236. [Google Scholar] [CrossRef]
  205. Zou, X.; Li, H.; Bünger, M.; Egund, N.; Lind, M.; Bünger, C. Bone Ingrowth Characteristics of Porous Tantalum and Carbon Fiber Interbody Devices: An Experimental Study in Pigs. Spine J. 2004, 4, 99–105. [Google Scholar] [CrossRef]
  206. Jones, G.; Groza, J.R.; Yamazaki, K.; Shoda, K. Plasma Activated Sintering (PAS) of Tungsten Powders. Mater. Manuf. Process 1994, 9, 1105–1114. [Google Scholar] [CrossRef]
  207. Lifland, M.I.; Okazaki, K. Properties of Titanium Dental Implants Produced by Electro-Discharge Compaction. Clin. Mater. 1994, 17, 203–209. [Google Scholar] [CrossRef]
  208. Fambri, L. Book Review: Biomaterials. An Introduction, Second Edition Edited by J. B. Park and R. S. Lakes Plenum Press, New York, 1992 ISBN 0-306-43992-1, 394 pp. J. Bioact. Compat. Polym. 1993, 8, 289–290. [Google Scholar] [CrossRef]
  209. Ganesan, D.; Sellamuthu, P.; Prashanth, K.G. Vacuum Hot Pressing of Oxide Dispersion Strengthened Ferritic Stainless Steels: Effect of al Addition on the Microstructure and Properties. J. Manuf. Mater. Process. 2020, 4, 93. [Google Scholar] [CrossRef]
  210. Ibrahim, A.; Zhang, F.; Otterstein, E.; Burkel, E. Processing of Porous Ti and Ti5Mn Foams by Spark Plasma Sintering. Mater. Des. 2011, 32, 146–153. [Google Scholar] [CrossRef]
  211. Kashimbetova, A.; Slámecka, K.; Díaz-De-La-Torre, S.; Méndez-García, J.C.; Hernández-Morales, B.; Piña-Barba, M.C.; Hui, D.; Celko, L.; Montufar, E.B. Pressure-Less Spark Plasma Sintering of 3D-Plotted Titanium Porous Structures. J. Mater. Res. Technol. 2023, 22, 2147–2157. [Google Scholar] [CrossRef]
  212. Luo, S.D.; Qian, M.; Ashraf Imam, M. Microwave Sintering of Titanium and Titanium Alloys. In Titanium Powder Metallurgy: Science, Technology and Applications; Butterworth-Heinemann: Oxford, UK, 2015; pp. 237–251. [Google Scholar] [CrossRef]
  213. Tang, C.Y.; Wong, C.T.; Zhang, L.N.; Choy, M.T.; Chow, T.W.; Chan, K.C.; Yue, T.M.; Chen, Q. In Situ Formation of Ti Alloy/TiC Porous Composites by Rapid Microwave Sintering of Ti6Al4V/MWCNTs Powder. J. Alloys Compd. 2013, 557, 67–72. [Google Scholar] [CrossRef]
  214. Delgado-Ruiz, R.; Romanos, G. Potential Causes of Titanium Particle and Ion Release in Implant Dentistry: A Systematic Review. Int. J. Mol. Sci. 2018, 19, 3585. [Google Scholar] [CrossRef] [PubMed]
  215. Zhang, F.; Otterstein, E.; Burkel, E. Spark Plasma Sintering, Microstructures, and Mechanical Properties of Macroporous Titanium Foams. Adv. Eng. Mater. 2010, 12, 863–872. [Google Scholar] [CrossRef]
  216. Schmidt, R.; Pilz, S.; Lindemann, I.; Damm, C.; Hufenbach, J.; Helth, A.; Geissler, D.; Henss, A.; Rohnke, M.; Calin, M.; et al. Powder Metallurgical Processing of Low Modulus β-Type Ti-45Nb to Bulk and Macro-Porous Compacts. Powder Technol. 2017, 322, 393–401. [Google Scholar] [CrossRef]
  217. Kon, M.; Hirakata, L.M.; Asaoka, K. Porous Ti-6Al-4V Alloy Fabricated by Spark Plasma Sintering for Biomimetic Surface Modification. J. Biomed. Mater. Res. B Appl. Biomater. 2004, 68, 88–93. [Google Scholar] [CrossRef]
  218. Ertorer, O.; Topping, T.D.; Li, Y.; Moss, W.; Lavernia, E.J. Nanostructured Ti Consolidated via Spark Plasma Sintering. Metall. Mater. Trans. A Phys. Metall. Mater. Sci. 2011, 42, 964–973. [Google Scholar] [CrossRef]
  219. Zadra, M.; Casari, F.; Girardini, L.; Molinari, A. Microstructure and Mechanical Properties of Cp-Titanium Produced by Spark Plasma Sintering. Powder Metall. 2008, 51, 59–65. [Google Scholar] [CrossRef]
  220. Prashanth, K.G.; Wang, Z. Additive Manufacturing: Alloy Design and Process Innovations. Materials 2020, 13, 542. [Google Scholar] [CrossRef]
  221. Sivaprasad, K.; Babu, N.R.; Prashanth, K.G. Additive Manufacturing and Allied Technologies. Trans. Indian Inst. Met. 2023, 76, 269. [Google Scholar] [CrossRef]
  222. Maurya, H.S.; Vikram, R.J.; Kosiba, K.; Juhani, K.; Sergejev, F.; Suwas, S.; Prashanth, K.G. Additive Manufacturing of CMCs with Bimodal Microstructure. J. Alloys Compd. 2022, 938, 168416. [Google Scholar] [CrossRef]
  223. Maurya, H.S.; Kosiba, K.; Juhani, K.; Sergejev, F.; Prashanth, K.G. Effect of Powder Bed Preheating on the Crack Formation and Microstructure in Ceramic Matrix Composites Fabricated by Laser Powder-Bed Fusion Process. Addit. Manuf. 2022, 58, 103013. [Google Scholar] [CrossRef]
  224. Maurya, H.S.; Jayaraj, J.; Vikram, R.J.; Juhani, K.; Sergejev, F.; Prashanth, K.G. Additive Manufacturing of TiC-Based Cermets: A Detailed Comparison with Spark Plasma Sintered Samples. J. Alloys Compd. 2023, 960, 170436. [Google Scholar] [CrossRef]
  225. Maurya, H.S.; Marczyk, J.; Juhani, K.; Sergejev, F.; Kumar, R.; Hussain, A.; Akhtar, F.; Hebda, M.; Prashanth, K.G. Binder Jetting 3D Printing of Green TiC-FeCr Based Cermets—Effect of Sintering Temperature and Systematic Comparison Study with Laser Powder Bed Fusion Fabricated Parts. Mater. Today Adv. 2025, 25, 100562. [Google Scholar] [CrossRef]
  226. Horn, T.J.; Harrysson, O.L. Overview of Current Additive Manufacturing Technologies and Selected Applications. Sci. Prog. 2012, 95, 255–282. [Google Scholar] [CrossRef]
  227. Kurtz, S.; Ong, K.; Lau, E.; Mowat, F.; Halpern, M. Projections of Primary and Revision Hip and Knee Arthroplasty in the United States from 2005 to 2030. J. Bone Jt. Surg. Am. 2007, 89, 780–785. [Google Scholar] [CrossRef]
  228. Prashanth, K.; Löber, L.; Klauss, H.-J.; Kühn, U.; Eckert, J. Characterization of 316L Steel Cellular Dodecahedron Structures Produced by Selective Laser Melting. Technologies 2016, 4, 34. [Google Scholar] [CrossRef]
  229. Baskaran, J.; Muthukannan, D.; Shukla, R.; Konda Gokuldoss, P. Manufacturability and Deformation Studies on a Novel Metallic Lattice Structure Fabricated by Selective Laser Melting. Vacuum 2024, 222, 113065. [Google Scholar] [CrossRef]
  230. Jagadeesh, B.; Duraiselvam, M.; Prashanth, K.G. Deformation Behavior of Metallic Lattice Structures with Symmetrical Gradients of Porosity Manufactured by Metal Additive Manufacturing. Vacuum 2023, 211, 111955. [Google Scholar] [CrossRef]
  231. Ge, J.; Huang, J.; Lei, Y.; O’Reilly, P.; Ahmed, M.; Zhang, C.; Yan, X.; Yin, S. Microstructural Features and Compressive Properties of SLM Ti6Al4V Lattice Structures. Surf. Coat. Technol. 2020, 403, 126419. [Google Scholar] [CrossRef]
  232. Gokuldoss, P.K.; Kolla, S.; Eckert, J. Additive Manufacturing Processes: Selective Laser Melting, Electron Beam Melting and Binder Jetting—Selection Guidelines. Materials 2017, 10, 672. [Google Scholar] [CrossRef]
  233. Singh, N.; Hameed, P.; Ummethala, R.; Manivasagam, G.; Prashanth, K.G.; Eckert, J. Selective Laser Manufacturing of Ti-Based Alloys and Composites: Impact of Process Parameters, Application Trends, and Future Prospects. Mater. Today Adv. 2020, 8, 100097. [Google Scholar] [CrossRef]
  234. Ahn, D.G. Directed Energy Deposition (DED) Process: State of the Art. Int. J. Precis. Eng. Manuf. Green Technol. 2021, 8, 703–742. [Google Scholar] [CrossRef]
  235. Scudino, S.; Unterdörfer, C.; Prashanth, K.G.; Attar, H.; Ellendt, N.; Uhlenwinkel, V.; Eckert, J. Additive Manufacturing of Cu-10Sn Bronze. Mater. Lett. 2015, 156, 202–204. [Google Scholar] [CrossRef]
  236. Suryawanshi, J.; Prashanth, K.G.; Ramamurty, U. Mechanical Behavior of Selective Laser Melted 316L Stainless Steel. Mater. Sci. Eng. A 2017, 696, 113–121. [Google Scholar] [CrossRef]
  237. Suryawanshi, J.; Prashanth, K.G.; Scudino, S.; Eckert, J.; Prakash, O.; Ramamurty, U. Simultaneous Enhancements of Strength and Toughness in an Al-12Si Alloy Synthesized Using Selective Laser Melting. Acta Mater. 2016, 115, 285–294. [Google Scholar] [CrossRef]
  238. DebRoy, T.; Wei, H.L.; Zuback, J.S.; Mukherjee, T.; Elmer, J.W.; Milewski, J.O.; Beese, A.M.; Wilson-Heid, A.; De, A.; Zhang, W. Additive Manufacturing of Metallic Components—Process, Structure and Properties. Prog. Mater. Sci. 2018, 92, 112–224. [Google Scholar] [CrossRef]
  239. Karimi, J.; Zhao, C.; Prashanth, K.G. Massive Transformation in Dual-Laser Powder Bed Fusion of Ti6Al4V Alloys. J. Manuf. Process 2024, 119, 282–292. [Google Scholar] [CrossRef]
  240. Karimi, J.; Antonov, M.; Kollo, L.; Prashanth, K.G. Role of Laser Remelting and Heat Treatment in Mechanical and Tribological Properties of Selective Laser Melted Ti6Al4V Alloy. J. Alloys Compd. 2022, 897, 163207. [Google Scholar] [CrossRef]
  241. Yu, S.; Chi, H.; Li, P.; Guo, B.; Yu, Z.; Xu, Z.; Liang, P.; Zhang, Z.; Guo, Y.; Ren, L. Interpenetrating Phases Composites Ti6Al4V/Zn as Partially Degradable Biomaterials to Improve Bone-Implant Properties. Addit. Manuf. 2024, 93, 104411. [Google Scholar] [CrossRef]
  242. Liu, S.; Shin, Y.C. Additive Manufacturing of Ti6Al4V Alloy: A Review. Mater. Des. 2019, 164, 107552. [Google Scholar] [CrossRef]
  243. Wysocki, B.; Maj, P.; Sitek, R.; Buhagiar, J.; Kurzydłowski, K.J.; Świeszkowski, W. Laser and Electron Beam Additive Manufacturing Methods of Fabricating Titanium Bone Implants. Appl. Sci. 2017, 7, 657. [Google Scholar] [CrossRef]
  244. Sabban, R.; Bahl, S.; Chatterjee, K.; Suwas, S. Globularization Using Heat Treatment in Additively Manufactured Ti-6Al-4V for High Strength and Toughness. Acta Mater. 2019, 162, 239–254. [Google Scholar] [CrossRef]
  245. Gupta, S.K.; Shahidsha, N.; Bahl, S.; Kedaria, D.; Singamneni, S.; Yarlagadda, P.K.D.V.; Suwas, S.; Chatterjee, K. Enhanced Biomechanical Performance of Additively Manufactured Ti-6Al-4V Bone Plates. J. Mech. Behav. Biomed. Mater. 2021, 119, 104552. [Google Scholar] [CrossRef] [PubMed]
  246. Karimi, J.; Suryanarayana, C.; Okulov, I.; Prashanth, K.G. Selective Laser Melting of Ti6Al4V: Effect of Laser Re-Melting. Mater. Sci. Eng. A 2021, 805, 140558. [Google Scholar] [CrossRef]
  247. Sridharan, N.; Chaudhary, A.; Nandwana, P.; Babu, S.S. Texture Evolution During Laser Direct Metal Deposition of Ti-6Al-4V. Jom 2016, 68, 772–777. [Google Scholar] [CrossRef]
  248. Xu, W.; Brandt, M.; Sun, S.; Elambasseril, J.; Liu, Q.; Latham, K.; Xia, K.; Qian, M. Additive Manufacturing of Strong and Ductile Ti–6Al–4V by Selective Laser Melting via in Situ Martensite Decomposition. Acta Mater. 2015, 85, 74–84. [Google Scholar] [CrossRef]
  249. Subramanian, S.; Yadav, M.K.; Jayaraj, J.; Yangyang, F.; Xi, L.; Prashanth, K.G. Microstructural Homogenization through Laser Remelting in an Additively Manufactured Ti–40Nb Sample from Elemental Feedstock Powders. J. Mater. Res. Technol. 2025, 38, 4305–4320. [Google Scholar] [CrossRef]
  250. Zaharin, H.A.; Rani, A.M.A.; Azam, F.I.; Ginta, T.L.; Sallih, N.; Ahmad, A.; Yunus, N.A.; Zulkifli, T.Z.A. Effect of Unit Cell Type and Pore Size on Porosity and Mechanical Behavior of Additively Manufactured Ti6Al4V Scaffolds. Materials 2018, 11, 2402. [Google Scholar] [CrossRef]
  251. Wang, Z.; Xie, M.; Li, Y.; Zhang, W.; Yang, C.; Kollo, L.; Eckert, J.; Prashanth, K.G. Premature Failure of an Additively Manufactured Material. NPG Asia Mater. 2020, 12, 30. [Google Scholar] [CrossRef]
  252. Zhang, X.Y.; Fang, G.; Zhou, J. Additively Manufactured Scaffolds for Bone Tissue Engineering and the Prediction of Their Mechanical Behavior: A Review. Materials 2017, 10, 50. [Google Scholar] [CrossRef]
  253. Shanghai Cooyer New Material Co., L. Medical Titanium Alloy Anticorrosive Material. Available online: https://www.openpr.com/news/1188745/hip-replacement-implants-market-opportunity-2018-2025-by-top-key-players-zimmer-biomet-depuy-synthes-stryker-smith-and-nephew-b-braun-melsungen-ag-corin-arthrex-inc-evolutis.html (accessed on 30 November 2025).
  254. Technology Predictions 2022; 3D Printing in Healthcare. Available online: https://futuredirections.ieee.org/2022/01/27/technology-predictions-2022-3d-printing-in-healthcare/ (accessed on 30 November 2025).
  255. Metal AM, F.M. 3D Printing/A.M. Celebrating Ten Years of Metal Additively Manufactured Hip Cups. Available online: https://www.metal-am.com/celebrating-ten-years-of-metal-additively-manufactured-hip-cups/ (accessed on 30 November 2025).
  256. Fang, C.; Cai, H.; Kuong, E.; Chui, E.; Siu, Y.C.; Ji, T.; Drstvenšek, I. Surgical Applications of Three-Dimensional Printing in the Pelvis and Acetabulum: From Models and Tools to Implants. Unfallchirurg 2019, 122, 278–285. [Google Scholar] [CrossRef] [PubMed]
  257. Zhu, Y.; Babazadeh-Naseri, A.; Dunbar, N.J.; Brake, M.R.W.; Zandiyeh, P.; Li, G.; Leardini, A.; Spazzoli, B.; Fregly, B.J. Finite Element Analysis of Screw Fixation Durability under Multiple Boundary and Loading Conditions for a Custom Pelvic Implant. Med. Eng. Phys. 2023, 111, 103930. [Google Scholar] [CrossRef] [PubMed]
  258. Wong, K.C.; Kumta, S.M.; Gee, N.V.L.; Demol, J. One-Step Reconstruction with a 3D-Printed, Biomechanically Evaluated Custom Implant after Complex Pelvic Tumor Resection. Comput. Aided Surg. 2015, 20, 14–23. [Google Scholar] [CrossRef] [PubMed]
  259. Broekhuis, D.; Boyle, R.; Karunaratne, S.; Chua, A.; Stalley, P. Custom Designed and 3D-Printed Titanium Pelvic Implants for Acetabular Reconstruction after Tumour Resection. HIP Int. 2023, 33, 905–915. [Google Scholar] [CrossRef]
  260. Hollander, D.A.; Von Walter, M.; Wirtz, T.; Sellei, R.; Schmidt-Rohlfing, B.; Paar, O.; Erli, H.J. Structural, Mechanical and in Vitro Characterization of Individually Structured Ti-6Al-4V Produced by Direct Laser Forming. Biomaterials 2006, 27, 955–963. [Google Scholar] [CrossRef]
  261. Lin, C.Y.; Wirtz, T.; LaMarca, F.; Hollister, S.J. Structural and Mechanical Evaluations of a Topology Optimized Titanium Interbody Fusion Cage Fabricated by Selective Laser Melting Process. J. Biomed. Mater. Res. A 2007, 83, 272–279. [Google Scholar] [CrossRef]
  262. ATEC’s IdentiTiTM-PC ATEC’s IdentiTi Posterior Curved Porous Ti Interbody Implants. Available online: https://www.londonspine.com/instrumented-lumbar-fusion-for-degenerative-spondylolisthesis/ (accessed on 30 November 2025).
  263. Laura Griffiths Stroke Patient Gets Life Back with 3D Printed Cranial Implant from EOS. Available online: https://www.protolabs.com/en-gb/resources/blog/3d-printing-and-customised-medical-implants/ (accessed on 30 November 2025).
  264. Murr, L.E. Open-Cellular Metal Implant Design and Fabrication for Biomechanical Compatibility with Bone Using Electron Beam Melting. J. Mech. Behav. Biomed. Mater. 2017, 76, 164–177. [Google Scholar] [CrossRef]
  265. Mazzoli, A.; Germani, M.; Raffaeli, R. Direct Fabrication through Electron Beam Melting Technology of Custom Cranial Implants Designed in a PHANToM-Based Haptic Environment. Mater. Des. 2009, 30, 3186–3192. [Google Scholar] [CrossRef]
  266. Zhao, X.; Li, K.; Bai, X. Research on Impact Resistance of 3D Printing Titanium Alloy Personalized Cranial Prosthesis. J. Integr. Technol. 2022, 11, 48–56. [Google Scholar] [CrossRef]
  267. Yan, R.; Luo, D.; Huang, H.; Li, R.; Yu, N.; Liu, C.; Hu, M.; Rong, Q. Electron Beam Melting in the Fabrication of Three-Dimensional Mesh Titanium Mandibular Prosthesis Scaffold. Sci. Rep. 2018, 8, 750. [Google Scholar] [CrossRef]
  268. Moiduddin, K.; Mian, S.H.; Umer, U.; Ahmed, N.; Alkhalefah, H.; Ameen, W. Reconstruction of Complex Zygomatic Bone Defects Using Mirroring Coupled with EBM Fabrication of Titanium Implant. Metals 2019, 9, 1250. [Google Scholar] [CrossRef]
  269. Yánez, A.; Cuadrado, A.; Martel, O.; Afonso, H.; Monopoli, D. Gyroid Porous Titanium Structures: A Versatile Solution to Be Used as Scaffolds in Bone Defect Reconstruction. Mater. Des. 2018, 140, 21–29. [Google Scholar] [CrossRef]
  270. Hashmi, A.W.; Mali, H.S.; Meena, A.; Saxena, K.K.; Ahmad, S.; Agrawal, M.K.; Sagbas, B.; Valerga Puerta, A.P.; Khan, M.I. A Comprehensive Review on Surface Post-Treatments for Freeform Surfaces of Bio-Implants. J. Mater. Res. Technol. 2023, 23, 4866–4908. [Google Scholar] [CrossRef]
  271. Maleki, E.; Bagherifard, S.; Bandini, M.; Guagliano, M. Surface Post-Treatments for Metal Additive Manufacturing: Progress, Challenges, and Opportunities. Addit. Manuf. 2021, 37, 101619. [Google Scholar] [CrossRef]
  272. Hashmi, A.W.; Mali, H.S.; Meena, A. A Comprehensive Review on Surface Quality Improvement Methods for Additively Manufactured Parts. Rapid. Prototyp. J. 2023, 29, 504–557. [Google Scholar] [CrossRef]
  273. Mruthunjaya, M.; Yogesha, K.B. A Review on Conventional and Thermal Assisted Machining of Titanium Based Alloy. Mater. Today Proc. 2021, 46, 8466–8472. [Google Scholar] [CrossRef]
  274. Das, S.; Kibria, G.; Doloi, B.; Bhattacharyya Editors, B. Materials Forming, Machining and Tribology Advances in Abrasive Based Machining and Finishing Processes; Springer: Berlin/Heidelberg, Germany, 2020. [Google Scholar]
  275. Acharya, S.; Panicker, A.G.; Gopal, V.; Dabas, S.S.; Manivasagam, G.; Suwas, S.; Chatterjee, K. Surface Mechanical Attrition Treatment of Low Modulus Ti-Nb-Ta-O Alloy for Orthopedic Applications. Mater. Sci. Eng. C 2020, 110, 110729. [Google Scholar] [CrossRef]
  276. Acharya, S.; Suwas, S.; Chatterjee, K. Review of Recent Developments in Surface Nanocrystallization of Metallic Biomaterials. Nanoscale 2021, 13, 2286–2301. [Google Scholar] [CrossRef]
  277. Liu, Y.J.; Wang, H.L.; Li, S.J.; Wang, S.G.; Wang, W.J.; Hou, W.T.; Hao, Y.L.; Yang, R.; Zhang, L.C. Compressive and Fatigue Behavior of Beta-Type Titanium Porous Structures Fabricated by Electron Beam Melting. Acta Mater. 2017, 126, 58–66. [Google Scholar] [CrossRef]
  278. Zadpoor, A.A. Mechanical Performance of Additively Manufactured Meta-Biomaterials. Acta Biomater. 2019, 85, 41–59. [Google Scholar] [CrossRef]
  279. Ran, Q.; Yang, W.; Hu, Y.; Shen, X.; Yu, Y.; Xiang, Y.; Cai, K. Osteogenesis of 3D Printed Porous Ti6Al4V Implants with Different Pore Sizes. J. Mech. Behav. Biomed. Mater. 2018, 84, 1–11. [Google Scholar] [CrossRef] [PubMed]
  280. Panzavolta, S.; Torricelli, P.; Amadori, S.; Parrilli, A.; Rubini, K.; Della Bella, E.; Fini, M.; Bigi, A. 3D Interconnected Porous Biomimetic Scaffolds: In Vitro Cell Response. J. Biomed. Mater. Res. A 2013, 101, 3560–3570. [Google Scholar] [CrossRef] [PubMed]
  281. Dou, C.; Zhang, M.; Ren, D.; Ji, H.; Yi, Z.; Wang, S.; Liu, Z.; Wang, Q.; Zheng, Y.; Zhang, Z.; et al. Bi-Continuous Mg-Ti Interpenetrating-Phase Composite as a Partially Degradable and Bioactive Implant Material. J. Mater. Sci. Technol. 2023, 146, 211–220. [Google Scholar] [CrossRef]
  282. Li, Y.; Jahr, H.; Lietaert, K.; Pavanram, P.; Yilmaz, A.; Fockaert, L.I.; Leeflang, M.A.; Pouran, B.; Gonzalez-Garcia, Y.; Weinans, H.; et al. Additively Manufactured Biodegradable Porous Iron. Acta Biomater. 2018, 77, 380–393. [Google Scholar] [CrossRef]
  283. Shi, Z.Z.; Yu, J.; Liu, X.F.; Zhang, H.J.; Zhang, D.W.; Yin, Y.X.; Wang, L.N. Effects of Ag, Cu or Ca Addition on Microstructure and Comprehensive Properties of Biodegradable Zn-0.8Mn Alloy. Mater. Sci. Eng. C 2019, 99, 969–978. [Google Scholar] [CrossRef]
  284. Rad, H.R.B.; Idris, M.H.; Kadir, M.R.A.; Farahany, S. Microstructure Analysis and Corrosion Behavior of Biodegradable Mg–Ca Implant Alloys. Mater. Des. 2012, 33, 88–97. [Google Scholar] [CrossRef]
  285. Yadav, M.K.; Shukla, R.; Xi, L.; Wang, Z.; Prashanth, K.G. Metallic Multimaterials Fabricated by Combining Additive Manufacturing and Powder Metallurgy. J. Compos. Sci. 2025, 9, 80. [Google Scholar] [CrossRef]
  286. Shukla, R.; Yadav, M.K.; Madruga, L.Y.C.; Jayaraj, J.; Popat, K.; Wang, Z.; Xi, L.; Prashanth, K.G. A Novel Ti-Eggshell-Based Composite Fabricated by Combined Additive Manufacturing-Powder Metallurgical Routes as Bioimplants. Ceram. Int. 2025, 51, 6281–6291. [Google Scholar] [CrossRef]
  287. Zhang, M.; Zhao, N.; Yu, Q.; Liu, Z.; Qu, R.; Zhang, J.; Li, S.; Ren, D.; Berto, F.; Zhang, Z.; et al. On the Damage Tolerance of 3-D Printed Mg-Ti Interpenetrating-Phase Composites with Bioinspired Architectures. Nat. Commun. 2022, 13, 3247. [Google Scholar] [CrossRef]
Figure 1. Diagram showing the basic requirements for implant materials regarding compatibility, mechanical properties, and manufacturing.
Figure 1. Diagram showing the basic requirements for implant materials regarding compatibility, mechanical properties, and manufacturing.
Jmmp 09 00401 g001
Figure 2. A schematic diagram illustrating the timeline of osseointegration of a metallic implant with respect to cellular events for a duration of 28 days. Reproduced with permission from [122].
Figure 2. A schematic diagram illustrating the timeline of osseointegration of a metallic implant with respect to cellular events for a duration of 28 days. Reproduced with permission from [122].
Jmmp 09 00401 g002
Figure 4. Schematic diagram illustrates the classification of the four types of bone tissue cells. Reproduced with permissions from [130].
Figure 4. Schematic diagram illustrates the classification of the four types of bone tissue cells. Reproduced with permissions from [130].
Jmmp 09 00401 g004
Figure 5. Schematic representations illustrating the following: (a) the allotropic forms of titanium—hexagonal close-packed (HCP) α-Ti, which is stable up to 882 °C, and body-centered cubic (BCC) β-Ti, stable at temperatures above this point; and (b) the formation of various phases and corresponding phase diagrams resulting from the addition of different alloying elements to titanium. Reproduced with permissions from [159].
Figure 5. Schematic representations illustrating the following: (a) the allotropic forms of titanium—hexagonal close-packed (HCP) α-Ti, which is stable up to 882 °C, and body-centered cubic (BCC) β-Ti, stable at temperatures above this point; and (b) the formation of various phases and corresponding phase diagrams resulting from the addition of different alloying elements to titanium. Reproduced with permissions from [159].
Jmmp 09 00401 g005
Figure 6. Graphical representation illustrating different Ti-based implants in the human body for biomedical applications.
Figure 6. Graphical representation illustrating different Ti-based implants in the human body for biomedical applications.
Jmmp 09 00401 g006
Figure 7. Porosity and pore size observed in porous titanium structures fabricated using different manufacturing techniques. Reproduced with permissions from [160].
Figure 7. Porosity and pore size observed in porous titanium structures fabricated using different manufacturing techniques. Reproduced with permissions from [160].
Jmmp 09 00401 g007
Figure 8. Classification of various processing techniques employed to fabricate porous Ti structures.
Figure 8. Classification of various processing techniques employed to fabricate porous Ti structures.
Jmmp 09 00401 g008
Figure 9. Schematic diagram illustrates the process of fabricating porous titanium using space holder technique.
Figure 9. Schematic diagram illustrates the process of fabricating porous titanium using space holder technique.
Jmmp 09 00401 g009
Figure 10. (a) Images of the porous titanium structures produced with varying spacer content (X = 0.55, 0.6, 0.65, 0.7, and 0.75, respectively). (b) Microstructure of the porous titanium structures with varied spacer content. (c) Compression behavior of the porous structures with varied porosity content and (d) tabular representation of the porosity content with varying spacer content. Reproduced with permissions from [167].
Figure 10. (a) Images of the porous titanium structures produced with varying spacer content (X = 0.55, 0.6, 0.65, 0.7, and 0.75, respectively). (b) Microstructure of the porous titanium structures with varied spacer content. (c) Compression behavior of the porous structures with varied porosity content and (d) tabular representation of the porosity content with varying spacer content. Reproduced with permissions from [167].
Jmmp 09 00401 g010
Figure 11. (a) Secondary electron image of Ti-powder, (b) secondary electron image of PU foam, (c) flowchart describing the process of porous Ti fabrication via enhanced polymeric sponge replication method, (d) secondary electron image of the porous Ti produced without ethanol in the solvent, and (e) secondary electron image of porous Ti produced with ethanol in the solvent. Reproduced with permissions from [186].
Figure 11. (a) Secondary electron image of Ti-powder, (b) secondary electron image of PU foam, (c) flowchart describing the process of porous Ti fabrication via enhanced polymeric sponge replication method, (d) secondary electron image of the porous Ti produced without ethanol in the solvent, and (e) secondary electron image of porous Ti produced with ethanol in the solvent. Reproduced with permissions from [186].
Jmmp 09 00401 g011
Figure 12. (a) Porous titanium samples prepared by the entangled metal wire technique: (b) magnified image of the marked region in Figure 12. (a,c) Compressive behavior of the porous titanium structures produced by the entangled metal wire technique. Reproduced with permissions from [191].
Figure 12. (a) Porous titanium samples prepared by the entangled metal wire technique: (b) magnified image of the marked region in Figure 12. (a,c) Compressive behavior of the porous titanium structures produced by the entangled metal wire technique. Reproduced with permissions from [191].
Jmmp 09 00401 g012
Figure 21. Bioinspired architecture comprises additively manufactured Ti6Al4V and pressureless infiltrated Mg. The developed Mg-Ti composite resembles bioinspired architecture showing (a) brick-and-mortar structure, (b) Bouligand structure and (c) crossed-lamellar structure. Reproduced with permission from [287].
Figure 21. Bioinspired architecture comprises additively manufactured Ti6Al4V and pressureless infiltrated Mg. The developed Mg-Ti composite resembles bioinspired architecture showing (a) brick-and-mortar structure, (b) Bouligand structure and (c) crossed-lamellar structure. Reproduced with permission from [287].
Jmmp 09 00401 g021
Figure 22. Processing step for designing cranio-maxillofacial implant by infiltrating wollastonite and hydroxyapatite-based nanoscale bioceramics inside titanium alloys with arbitrary lattice structures. Reproduced with permissions from [27].
Figure 22. Processing step for designing cranio-maxillofacial implant by infiltrating wollastonite and hydroxyapatite-based nanoscale bioceramics inside titanium alloys with arbitrary lattice structures. Reproduced with permissions from [27].
Jmmp 09 00401 g022
Table 1. Physical and mechanical properties of metallic, ceramic, and polymeric biomaterials (ρ—density in g/cm3; E—Young’s modulus in GPa; YS—yield strength in MPa; UTS—ultimate tensile strength in MPa; UCS—Ultimate compressive strength in MPa; and FS—fatigue strength in MPa, 107 cycles).
Table 1. Physical and mechanical properties of metallic, ceramic, and polymeric biomaterials (ρ—density in g/cm3; E—Young’s modulus in GPa; YS—yield strength in MPa; UTS—ultimate tensile strength in MPa; UCS—Ultimate compressive strength in MPa; and FS—fatigue strength in MPa, 107 cycles).
MaterialρEYSUTSUCSFSRef.
Natural Bone
Cortical bone1.8–2.07–30-164–240100–23027–35[68]
Cancellous bone1.0–1.40.01–3.0--2–12-
Metals and Alloys
Ti-6Al-4V (cast)4.43114760–880895–930-600–700[68]
Ti-6Al-4V (wrought)4.43114827–1103860–965896–1172500–800
Ti-6Al-7Nb4.52105880900--[15]
SS316L8.0193170–310540–1000480–620240–480[68]
Fe20Mn7.73207420700--[69]
Zn-Al-Cu5.7990171210--[34]
Co-Cr-Mo8.3240500–1500900–1540-500–900
CoCr20Ni15Mo77.8195–230240–450450–960--
Pure Mg (cast)1.7441218740-
Pure Mg (wrought)1.7441100180100–140-
AZ31 (Mg-based alloy)1.7845185263--
AZ91 (Mg-based alloy)1.8145160150--
Ceramics
Alumina Ceramics4260–410-400–580--[34]
Synthetic hydroxyapatite3.156–102--0.22–4.1-[18,70]
Zirconia3.98210-800–15001990-[71]
Polymers
PLGA1.2–1.31.693.8–26.613.9–16.7--[34]
PCL1.15281–6868.37–14.6668–103--
PLA1.837507059--
Table 3. List showing the mechanical properties of titanium and its alloys (E—modulus in GPa; UTS—ultimate tensile strength in MPa) [15].
Table 3. List showing the mechanical properties of titanium and its alloys (E—modulus in GPa; UTS—ultimate tensile strength in MPa) [15].
MaterialStandardEUTSAlloy Composition
First-generation biomaterials (1950–1990)
Commercially pure Ti (CP grade 1–4)ASTM F1341100240–550α
Ti–6Al–4V ELI wroughtASTM F136110860–965α + β
Ti–6Al–4V ELI standard gradeASTM F1472112895–930α + β
Ti–6Al–7Nb wroughtASTM F1295110900–1050α + β
Ti–5Al–2.5Fe-1101020α + β
Second-generation biomaterials (1990–to date)
Ti–13Nb–13Zr wroughtASTM F171379–84973–1037Metastable β
Ti–12Mo–6Zr–2Fe (TMZF)ASTM F181374–851060–1100β
Ti–35Nb–7Zr–5Ta (TNZT)-55596β
Ti–29Nb–13Ta–4.6Zr-65911β
Ti–35Nb–5Ta–7Zr–0.40 (TNZTO)-661010β
Ti–15Mo–5Zr–3Al-22-β
Ti–MoASTM F2066--β
Table 4. Mechanical properties of titanium scaffolds prepared by different space holders (P—porosity in %; PS—pore size (µm); E—Young’s modulus in GPa; YS—Yield strength in MPa; UTS—ultimate tensile strength in MPa; UCS—ultimate compressive strength in MPa).
Table 4. Mechanical properties of titanium scaffolds prepared by different space holders (P—porosity in %; PS—pore size (µm); E—Young’s modulus in GPa; YS—Yield strength in MPa; UTS—ultimate tensile strength in MPa; UCS—ultimate compressive strength in MPa).
Space Holder MaterialPPSEYSUTSUCSRef.
Mo Wire32–47-23–6276–192--[175]
Mg45–705250.42–8.815–116--[164]
Mg50–71262–132---59–280[163]
Mg30–50-15–44117–222--[176]
Ti Fibers35–84150–6002–4-200–600-[177]
Rice Husk50–60100–550---17–70[165]
Rice Husk25–36----440–938[166]
Rice Husk15–34-6–15--116–396[16]
Sucrose20–54212–50012–50---[173]
Urea36480----[168]
Urea55–75200–5003–6-10–35-[169]
Table 5. Mechanical properties of porous titanium prepared by entangled metal wire technique (P—porosity in %; PS—pore size (µm); E—Young’s modulus in GPa; YS—yield strength in MPa; UTS—ultimate tensile strength in MPa; UCS—ultimate compressive strength in MPa; FS—flexural strength in MPa; NR— not reported).
Table 5. Mechanical properties of porous titanium prepared by entangled metal wire technique (P—porosity in %; PS—pore size (µm); E—Young’s modulus in GPa; YS—yield strength in MPa; UTS—ultimate tensile strength in MPa; UCS—ultimate compressive strength in MPa; FS—flexural strength in MPa; NR— not reported).
MethodMaterial UsedPPSEYSUTSUCSFSRef.
EWMTEntangled Mo Wire32–470.423–6276–192---[175]
EWMTTi Wire35–84150–6002–4.2-200–600--[177]
EWMTEntangled Ti Wires44–81NR0.03–2.25---9–325[193]
EWMTEntangled Ti Wire53–55NR0.03–13–3.5---[188]
EWMTEntangled Ti Wire37–54NR22–47--175–246-[191]
EWMTEntangled Ti Wire40–55100–4000.4–1.412.9–52.5---[192]
EWMTEntangled Ti Wire45–5850–2001.05–0.3375–12448–108--[187]
EMWTNormally entangled Ti Wire48–73-0.13–0.822–31---[190]
Coiled entangled Ti Wire48–78-0.04–0.621–19---
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

Yadav, M.K.; Yarlapati, A.; Aditya, Y.N.; Kesavan, P.; Pandey, V.; Perugu, C.S.; Nain, A.; Chatterjee, K.; Suwas, S.; Jayamani, J.; et al. Processing and Development of Porous Titanium for Biomedical Applications: A Comprehensive Review. J. Manuf. Mater. Process. 2025, 9, 401. https://doi.org/10.3390/jmmp9120401

AMA Style

Yadav MK, Yarlapati A, Aditya YN, Kesavan P, Pandey V, Perugu CS, Nain A, Chatterjee K, Suwas S, Jayamani J, et al. Processing and Development of Porous Titanium for Biomedical Applications: A Comprehensive Review. Journal of Manufacturing and Materials Processing. 2025; 9(12):401. https://doi.org/10.3390/jmmp9120401

Chicago/Turabian Style

Yadav, Mayank Kumar, Akshay Yarlapati, Yarlapati Naga Aditya, Praveenkumar Kesavan, Vaibhav Pandey, Chandra Shekhar Perugu, Amit Nain, Kaushik Chatterjee, Satyam Suwas, Jayaraj Jayamani, and et al. 2025. "Processing and Development of Porous Titanium for Biomedical Applications: A Comprehensive Review" Journal of Manufacturing and Materials Processing 9, no. 12: 401. https://doi.org/10.3390/jmmp9120401

APA Style

Yadav, M. K., Yarlapati, A., Aditya, Y. N., Kesavan, P., Pandey, V., Perugu, C. S., Nain, A., Chatterjee, K., Suwas, S., Jayamani, J., & Konda Gokuldoss, P. (2025). Processing and Development of Porous Titanium for Biomedical Applications: A Comprehensive Review. Journal of Manufacturing and Materials Processing, 9(12), 401. https://doi.org/10.3390/jmmp9120401

Article Metrics

Back to TopTop