Download as pdf or txt
Download as pdf or txt
You are on page 1of 49

Progress in Materials Science 88 (2017) 232–280

Contents lists available at ScienceDirect

Progress in Materials Science


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

Advances in metals and alloys for joint replacement


Lenka Kunčická a, Radim Kocich a, Terry C. Lowe b,⇑
a
Regional Materials Science and Technology Centre, VSB – Technical University of Ostrava, 17. Listopadu 15, 70833 Ostrava-Poruba, Czech Republic
b
The George S. Ansell Department of Metallurgical and Materials Engineering, Colorado School of Mines, 1500 Illinois Street, Golden, CO 80401, USA

a r t i c l e i n f o a b s t r a c t

Article history: Metals, ceramics, polymers, and composites have been employed in joint arthroplasty with
Received 12 March 2016 ever increasing success since the 1960s. New materials to repair or replace human skeletal
Received in revised form 22 December 2016 joints (e.g. hip, knee, shoulder, ankle, fingers) are being introduced as materials scientists
Accepted 3 April 2017
and engineers develop better understanding of the limitations of current joint replacement
Available online 8 April 2017
technologies. Advances in the processing and properties of all classes of materials are pro-
viding superior solutions for human health. However, as the average age of patients for
Keywords:
joint replacement surgery decreases and the average lifespans of men and women
Arthroplasty
Joint replacement
increases worldwide, the demands upon the joint materials are growing. This article
Medical alloys focuses solely on advances in metals, highlighting the current and emerging technologies
Review in metals processing, metal surface treatment, and integration of metals into hybrid mate-
Surface treatment rials systems. The needed improvements in key properties such as wear, corrosion, and
fatigue resistance are discussed in terms of the enhanced microstructures that can be
achieved through advanced surface and bulk metal treatments. Finally, far reaching hori-
zons in metals science that may further increase the effectiveness of total joint replace-
ment solutions are outlined.
Ó 2017 Elsevier Ltd. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 233
2. Design of joint replacement materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 234

Abbreviations: 3DP, 3D Printing; ALM, additive layer manufacturing; AM, additive manufacturing; ARB, accumulative roll bonding; ASTM, American
Society of Testing Materials; AWJ, abrasive water jet; BCC, body centered cubic; CE, Conformité Européene; CP Ti, commercial purity titanium; CT,
computed tomography; DLF, directed light fabrication; DMLS, direct metal laser sintering; EBFF, electron beam freeform fabrication; EBM, electron beam
melting; EDM, electro discharge machining; EMAT, electromagnetic acoustic transducer; FCC, face centered cubic; FDA, Food and Drug Administration; FEA,
finite element analysis; FEM, finite element method; FGM, functionally graded materials; HA, hydroxyapatite; HCP, hexagonal close packed; HPT, high
pressure torsion; HR-pQCT, high resolution peripheral quantitative computed tomography; LBMD, laser based metal deposition; LENSTM, laser engineered
net shaping; LFIT, low friction ion treated; LEHT, low energy-high temperature ion implantation; LSEM, large strain extrusion machining; MAM, modulation
assisted machining; PIII, plasma immersion ion implantation; PSII, plasma source ion implantation; PCL, poly (caprolactone); PE, poly (ethylene); PGA, poly
(glycolic acid); PHB, poly (hydroxybutyrate); PIRAC, powder immersion reaction assisted coating; PMMA, poly (methyl methacrylate); PM, powder
metallurgy; RP, rapid prototyping; RUS, resonant ultrasound spectroscopy; SFE, stacking fault energy; SFF, solid freeform fabrication; SLM, selective laser
melting; SLS, selective laser sintering; SPD, severe plastic deformation; THA, total hip arthroplasty; TKA, total knee arthroplasty; TMZF, titanium
molybdenum zirconium iron; TNTZ, titanium niobium tantalum zirconium; UNS, unified numbering system; UTS, ultimate tensile strength; UCS, ultimate
compressive strength; UFG, ultrafine grain; UHMWPE, ultra-high molecular weight polyethylene; VAR, vacuum arc remelting; YS, yield strength.
⇑ Corresponding author.
E-mail address: lowe@mines.edu (T.C. Lowe).

http://dx.doi.org/10.1016/j.pmatsci.2017.04.002
0079-6425/Ó 2017 Elsevier Ltd. All rights reserved.
L. Kunčická et al. / Progress in Materials Science 88 (2017) 232–280 233

2.1. State of art . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 235


2.2. Computer simulations. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 237
3. Current metals and alloys and their microstructures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 237
3.1. Stainless steels. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 237
3.2. Cobalt-chromium alloys . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 238
3.3. Titanium and its alloys . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 239
3.4. Tantalum alloys . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 241
3.5. Zirconium alloys . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 241
4. Mechanical properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 243
4.1. Strength . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 243
4.2. Elastic stiffness . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 246
4.3. Wear. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 247
4.4. Fatigue . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 249
4.5. Corrosion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 249
4.6. Compatibility with non-metallic materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 251
4.7. Biological properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 251
5. Current treatment of joint alloy surfaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 252
5.1. Treatments to reduce wear. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 253
5.2. Treatments to enhance fatigue strength . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 254
5.3. Treatments to enhance bone tissue integration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 255
6. Advances in metals processing for joint replacements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 256
6.1. Advances in primary metal processing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 256
6.2. Advanced wrought metals processing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 258
6.3. Severe plastic deformation processes. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 259
6.4. Particle based metals fabrication . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 260
6.5. Additive manufacturing processes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 260
7. Advances in treating metallic joint surfaces. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 262
7.1. Sub-surface modification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 264
7.2. Sub-surface addition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 265
7.2.1. Ion implantation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 265
7.2.2. Thermochemical diffusion methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 266
7.3. Top-surface modification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 267
7.4. Top-surface addition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 267
8. Outlooks and horizons . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 268
8.1. Horizons in materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 269
8.2. Horizons in technologies. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 269
9. Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 270
Acknowledgement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 270
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 270

1. Introduction

High performance materials have been one of the central enablers of joint replacement technology. While ceramics, poly-
mers, and metals have all been used in prosthetic joints, the focus of this work is on the longest enduring and most widely
used joint replacement material class: metals and their alloys. This paper examines the status of metals and alloys in current
joint replacement systems, as well as identifies the horizons of alloy technology that may appear in new orthopedic device
offerings in the coming years. These systems include orthotic and prosthetic appliances most commonly used to augment or
replace the function of the ankle, knee, hip, spine, shoulder, elbow, and fingers.
One of the most prominent demands that joint arthroplasty imposes on joint materials is the need for them to survive
relative motion via sliding and highly localized loads. Thus, wear and surface properties of joint replacement alloys are of
utmost importance. In Section 2, the history of joint replacement is reviewed, including the dependence of the evolution
of joint design on advancements in metals technology. The mechanical, physical, and biological properties that are most
important in joint replacement applications of metals and alloys are highlighted in Section 3. In Section 4, the major families
of alloys and the key characteristics of their microstructures that make them useful in current joint replacement systems are
systematically summarized. Section 5 focuses on current surface treatment and modification technologies, while Sections 6
and 7 then address emerging advances in alloy and manufacturing and finishing technologies that can enhance the perfor-
mance and increase the longevity of joint replacements. Combining insights from Sections 6 and 7, Section 8 then identifies
likely horizons for materials technology that may, in the long term, result in a completely new generation of joint replace-
234 L. Kunčická et al. / Progress in Materials Science 88 (2017) 232–280

ments. Finally, Section 9 concludes with a synopsis of current, short term, and long term joint replacement alloy
technologies.
Many excellent reviews and original articles that highlight or summarize aspects of metals and alloys for biomedical
applications, including joint replacement already exist. However, throughout this review there is a persistent emphasis
on processing methods of metals and their resulting microstructures [1–11]. This work will show that the most significant
foundations for the evolution of joint replacement alloys are being provided by advances in bulk metal and surface process-
ing methods, plus other novel manufacturing technologies.

2. Design of joint replacement materials

Biocompatible materials are generally used to replace or support damaged biological tissues. They can be applied to
replace hard tissues in orthopedy or orthodontia, or to support soft tissues, e.g. as blood vessel stents or artificial valves
in the heart [12,13]. The spine, hip and knee are the body parts which most often need implants due to influences of degen-
erative diseases, such as osteoporosis, osteoarthritis, or mechanical damage. However, non-biological replacements
implanted into living organisms always act as foreign bodies and therefore the materials have to be adaptable to the sur-
rounding environment within the organisms [14]. They should feature excellent biocompatibility and corrosion resistance,
good fatigue and wear resistance and the combination of strength and elastic modulus should correspond as much as pos-
sible to the properties of human bones. Last but not least, the demand for greater longevity of implants is ever increasing in
order to prevent revision surgery, which is expensive and can be painful for the patient [15]. Increasing the longevity of pri-
mary implants is an important issue also from the viewpoint of increasing human life expectancy. All the above factors have
spurred continuous development of new biocompatible materials and enhancement of properties of current materials.
The history of human joint replacements goes back to the 19th century; the first ever implantation of a large human joint
was performed in 1890 [16]. However, the first implants were made from natural materials. The first total knee endopros-
thesis was made of ivory and anchored to the bones using nickel-coated nails and a mixture of gypsum, resin and pumice.
Later on, materials based on polymers and glasses were introduced [15]. The first ivory and glass hip joint was implanted in
1926, while implantation of metal-based joint prostheses began in the early 1930s.
The first ever stainless steel total hip replacement was implanted in 1938. It was made of the ‘‘vanadium steel” purposely
designed to be used within a human organism [17]. However, this type of steel had inadequate biocompatibility and corro-
sion resistance. Consequently, beginning in the 1950s superior alloys such as 316L stainless steel were developed. Together
with introduction of stainless steel implants, the development of cobalt alloys began [15]. The first Co-based alloy implant
was introduced in 1939, while the first application of the castable CoCrMo alloy known at present as VitalliumÒ was in the
1940s in dentistry. Later on, this alloy was used for fabrication of artificial joint replacements owing to its superior mechan-
ical properties and corrosion resistance when compared to other materials used at that time [18,19]. The first successful
implantation of a metal-on-metal hip joint replacement (McKee design), fabricated from steel and later from CoCr alloys,
was performed in the 1950s [20]. Nevertheless, the metal-on-metal implants exhibited tendencies to loosen from bone
despite continuous development and numerous technological improvements. The longevity of implants increased signifi-
cantly with the development of acrylic bone cement (polymethylmethacrylate – PMMA); the enhanced McKee-Farrar
metal-on-metal hip joint replacement featuring bone cement fixation exhibited a relatively long durability and stability dur-
ing the 1960s [15].
Independent of the two above mentioned material groups, development of Ti-based alloys suitable for fabrication of joint
implants started in the 1930s [20]. After World War II, two Ti materials – commercially pure Ti (CP Ti) and Ti6Al4V – gained
the dominant position, although the possibility of medical use of Ti-based shape memory alloys (SMA) was investigated con-
currently during the 20th century [21,22]. SMAs were discovered in the early 1960s and finally introduced into biomechanics
in 1973 [15]. Although they featured very low elastic modulus ranging between 30 and 50 MPa while possessing strength
comparable to stainless steels, clinical orthopedic studies performed in the 1970s and 1980s revealed poor fatigue proper-
ties, which finally prevented them from being used for load-bearing joint replacements components. Nevertheless, SMA
alloys are advantageously used for example as orthodontic equipment and temporary spine and long bones fixations [23].
Alloying of titanium with elements such as zirconium, niobium, molybdenum, tin and tantalum led to the development
of b phase Ti alloys, which appeared to have properties favorable for biomedical applications. These alloys were introduced
in the 1990s and their continuous improvement has been one of the leading topics of orthopedic materials research since
then.
Besides being an alloying element in b phase Ti alloys, tantalum has been used for production of auxiliary equipment for
surgery of nerves and vessels since the 1950s. However, once the properties of tantalum had been explored in depth (espe-
cially when its superior corrosion resistance was revealed), the efforts to produce a ‘‘perfect” implant were no longer limited
only to the Ti, CoCr, and stainless steel alloys. The present use of tantalum is mostly for porous coatings, which were recently
developed to increase implant longevity. Tantalum coatings favorably influence tribology of individual implant components,
as well as their biocompatibility.
Since the 1970s non-metal-based materials for artificial joint components evolved in parallel with metal-based bio-
applicable materials. The first completely ceramic joint replacement (Al2O3) was implanted in 1970 [20]. Although this
design featured a hard contact, the wear rate was significantly lower than for other materials available at the time. To
L. Kunčická et al. / Progress in Materials Science 88 (2017) 232–280 235

Fig. 1. Historical milestones in the development of joint replacements.

generally decrease the wear rate, orthopedic surgeon Sir John Charnley introduced the ‘‘low-friction arthroplasty” concept
for hip joint implants [24]. This at first consisted of covering the acetabular and femoral components with polytetrafluo-
rethylene (PTFE) shells, which was not successful and exhibited significant debonding and wear [20]. Soon after, ultra-
high molecular weight polyethylene (UHMWPE) was introduced [14]. Nevertheless, UHMWPE suffered from high wear in
the metal-on-polymer contacts of joint components. After these not entirely successful efforts, metal-on-metal joint replace-
ment designs were revived in the 1980s and the research expanded to include improvement of other materials, which
resulted in the introduction of zirconia oxide ceramics in 1985. Replacement of Al2O3 ceramics by zirconia led to improve-
ment of mechanical properties and wear resistance of hip joint femoral heads [20].

2.1. State of art

At present, the most widely used metals for joint replacements are stainless steels, cobalt-chromium alloys and titanium
alloys. These three main alloy groups have a relatively long tradition in joint replacements (documented by Fig. 1 depicting
historical milestones throughout the history of their development), and remain favorites due to their good biocompatibility
and the ability to undergo large plastic deformations prior to failure [15]. Nevertheless, limiting factors still exist especially
for steels and Co-based alloys. Significant problems include wear and fretting corrosion fatigue of hip joint implants acetab-
ular components, both of which can lead to loosening of the implant. However, such problems can be eliminated by devel-
opment of new materials with enhanced properties or improvement of existing ones.
Stainless steel is a general label for iron-based alloys featuring high contents of chromium and substantial amounts of
nickel [15]. The 316L type stainless steel has been widely used for applications such as fracture plates, screws and nails. Cer-
tain joint replacement components, such as femoral and tibial components of knee prostheses, are typically fabricated from
this steel as well [25]. However, the trend is to replace stainless steels in permanent implants components with new mate-
rials featuring properties more like human bones. On the other hand, use of steel for temporary applications, such as fracture
fixations, is quite common and favorable. Nickel is problematic due to its toxicity and overall harmful effect on living
236 L. Kunčická et al. / Progress in Materials Science 88 (2017) 232–280

organisms. Therefore, nickel-free stainless steels were developed, resulting in the introduction of steels with increased nitro-
gen content (nickel-free high nitrogen steels – HNS) [26]. Since then, HNS have gained the attention of many commercial
companies such as Carpenter Technology Corp., which developed the BioDur 108 steel (ASTM F2229) with enhanced
mechanical properties and resistance to corrosion. HNS (e.g. BIOSSN4 and P558) in general exhibit good overall biocompat-
ibility [23,24]. The Orthinox steel, which is at present commonly used for fabrication of joint implant components such as
stems of total hip replacements, was developed in 1998 [15].
Considering Co-based alloys, CoCrMo stems are less prone to fatigue cracks due to corrosion when compared to steel
stems [15]. Even though the majority of total hip joint replacement stems are still fabricated from steels, especially from
those with enhanced properties (Orthinox), approximately 20% of the entire production of this component is presently made
from wrought CoCrMo. Femoral and tibial components of knee implants are also usually fabricated from CoCrMo alloys (and
typically equipped with UHMWPE linings). Modifications of CoCrMo are the primary choice for the majority of knee and
ankle joint replacements [25]. CoCr alloys with additions of W and Ni are particularly used for temporary and short-term
implants, while stems of heavily loaded joints, such as knee and hip, are fabricated from more recently developed CoNiCrMo
alloys [15]. Nevertheless, the main trend for these alloys is not to design new chemical compositions, but to improve
mechanical properties of the existing alloys via implementing new processing technologies and thermo-mechanical treat-
ments [20]. For example, a new fabrication technology for the classic CoCrMo ATSM F75 alloy ensuring enhanced wear resis-
tance and fatigue strength has recently been patented (TJA-1537).
Due to their excellent mechanical properties approaching the properties of human bones (high strength, low density, rel-
atively low elastic modulus) and high biocompatibility, specific strength and corrosion resistance, titanium alloys began to
be widely used for joint replacements together with the above-mentioned materials [20]. Nowadays, the most commonly
used Ti-based biomaterials are Commercial Purity (CP) Ti (ASTM F67) and Ti6Al4V (ASTM F136) alloy. CP Ti and Ti6Al4V
ELI (extra low interstitial) are typically used in dental and spinal surgery, while Ti6Al4V is used to fabricate components
of artificial knee, hip and shoulder joints, as well as for bone fixators. Due to the toxicity of vanadium, various vanadium-
free Ti6Al4V modifications have been introduced for applications such as femoral stems, spinal components and fracture fix-
ation plates. Nevertheless, a significant focus of recent research is the development of b phase Ti alloys with properties that
are even more comparable to the properties of human bones, and are moreover also usually free of vanadium. A typical
example is the TiMoZrFe alloy, already practically applied for hip joint stems [15]. Unfortunately, even the newly introduced
b phase alloys still feature higher elastic moduli than human bones. The efforts are therefore not only to modify chemical
compositions, but also to design innovative fabrication methods and post-production treatments. Among the research trends
leading to development of prostheses featuring minimum differences between their mechanical properties and the proper-
ties of bones are new casting methods, innovative powder metallurgy fabrication processes for porous structures develop-
ment, progressive treatments including thermo-mechanical processing and, last but not least, modification of structure via
severe plastic deformation (SPD) technologies [27–29]. However, production processes usually improve one type of proper-
ties at the expense of another and thus methods to influence the properties separately without harming others have been
researched.
One of the favorable ways to selectively enhance properties is the application of surface coatings. For example, compo-
nents made from CoCrMo alloys exhibit high strength and durability with poorer wear resistance, which is problematic espe-
cially for components experiencing high loads and hard contacts, such as heads of hip joint implants. The durability and
longevity of the entire implant can be significantly increased by application of a coating [30]. Surface coatings can be based
on various materials, including metals. Among the coatings with the best performance are the recently introduced highly
durable and wear resistant tantalum-based porous coatings, such as Trabecular MetalTM (see Section 3.4) [31]. These coatings
feature macroporous structure advantageously supporting osseointegration. Tantalum is not only used for coatings, but also
for individual Ta-based alloys, especially in combinations with titanium and additions of carbon, which are already applied
as spinal implants and vertebral members replacements [31,32]. In orthopedy, Ta-based materials are promising especially
due to their potential to increase the longevity of knee and hip joint implants [15]. However, the longevity of joint replace-
ments can also be increased by covering of their components with durable ceramics. One ceramic frequently used for
femoral heads and acetabular liners is zirconium oxide [26]. Although ceramics are relatively brittle compared to metals,
they typically feature a very hard wear resistant surface. Despite the relatively benign behavior of ceramics within physio-
logical environments, their rather inflexible combination of characteristics makes them difficult to handle through the joint
replacement fabrication process. The first ceramic-on-ceramic hip joint replacement design was approved by United States
FDA (Food and Drug Administration) just in 2003, although it had already been used in Europe at that time [20]. Ceramic
materials are also typical components of functionally graded materials (FGM), which are promising implant materials due
to their graduated chemical composition, structure and properties [33].
Zirconium is attractive not only in bioceramics, but also in zirconium alloys. ZrNb alloys were recently introduced into
production of knee implant components, the longevity of which is predicted to be 20–25 years [15]. Modern joint replace-
ment designs also typically contain other ceramic and polymeric components. Among the most innovative Zr-based mate-
rials for knee and hip replacements is OxiniumTM, oxidized zirconium developed by Smith and Nephew Inc., introduced in
1997 into knee implants and in 2002 into hip implants [34]. This material basically consists of a Zr metal alloy, the surface
of which has been transformed to ceramics via a patented oxidation process. OxiniumTM combines high strength and ductility
typical of metal alloys with the advantages of ceramic coatings – very hard surface layer and superior resistance to friction
and abrasion.
L. Kunčická et al. / Progress in Materials Science 88 (2017) 232–280 237

Among the above mentioned, a variety of other materials is also used in biomedicine. One of the trends in the research
and development of biocompatible metal-based materials are bio-absorbable Mg alloys for temporary bone fixations [35],
however, these materials are not suitable for joint implants due to their low mechanical properties. Nevertheless, many
detailed studies dealing with stents and other implants and contemporary fixations have been published.

2.2. Computer simulations

Development of new innovative materials and replacement designs demands thorough testing of their behavior in
extreme conditions, such as pH changes, high compressive forces, localized pressure and cyclic loading. Although certain
properties (e.g. corrosion behavior) can be tested in chemical laboratories, others, such as contact stress and its distribution
within a joint or determination of compressive forces, can only hardly be measured experimentally since the actions of
motions and loads in a joint during movement (i.e. at different flexion angles) are very complex.
Computer simulations are useful for predicting mechanical conditions within joint replacements, and are finding increas-
ing use in biomedicine and biomechanics. Numerical predictions via the finite element method (FEM) can be used to perform
various engineering analyses [36–38], including predictions of mechanical behavior (fatigue life, locations of failure, etc.) of
joints and implants in orthopedics [39–43]. Simulations can also aid the development of new technologies for processing of
implant materials [44,45], design of new models of replacements [46], and in evaluation of characteristics of various implant
designs [47]. Comprehensive studies of healthy joints are also important, since their detailed numerical models can subse-
quently be used to optimize models of artificial joints and to determine the boundary and external conditions [48]. FE anal-
ysis is therefore a time-saving and cost-saving tool for development of new materials and designs in biomedicine. However,
to carry out a successful FE simulation, the input model must be assembled with the greatest attention. Models of bones,
onto which the implant components are attached, are typically based on CT (computed tomography) scans [46], while
the input parameters necessary to define materials of the individual implant components are usually acquired experimen-
tally via mechanical testing (of loosened prostheses) [49]. Nevertheless, the primary condition for effective testing (exper-
imental or numerical) is to also have detailed descriptions of the constituent materials, since only then can the
experimental and/or predicted results be well understood and correctly interpreted. The following sections therefore deal
with the basic descriptions of microstructures (Section 3) and mechanical properties (Section 4) of the most commonly used
replacement materials.

3. Current metals and alloys and their microstructures

3.1. Stainless steels

In general, stainless steels can be characterized by their chemical compositions (chromium type and chromium-nickel
type steels) and their microstructures [15]. The microstructures can be either ferritic, the lattices of which are body centered
cubic (BCC), martensitic, with tetragonal lattices, or austenitic, with face centered cubic (FCC) lattice structures. The latter
two specifically find their application in medicine. Martensitic steels are used to produce surgical and dental instruments,
while austenitic steels are the only steel type suitable for permanent and long-term surgical implants and joint replacements
since their FCC crystal structure makes them non-ferromagnetic. However, they are also widely used for auxiliary medical
non-implantable equipment and devices due to their good corrosion resistance [26].

Fig. 2. Microstructures depicting twins in (a) low-C high-Cr, Ni 316LN stainless steel; (b) Ni-free high-Mn 317 L nitrogen stainless steel [25,380].
238 L. Kunčická et al. / Progress in Materials Science 88 (2017) 232–280

Stainless steels used within humans must meet high quality standards. They must possess a high level of cleanliness and
minimum impurity content, which is nowadays typically achieved by innovative production technologies, such as vacuum
melting, vacuum arc re-melting, and electro-slag refining [15]. For example, the most widely used stainless steel, 316 (ATSM
F138), can have its designation supplemented with the letters ‘‘L” and ‘‘V”, where ‘‘L” stands for a low carbon content, while
the vacuum under which the steel was melted is identified by ‘‘V” [20].
All austenitic stainless steels contain 10.5–20 wt.% of chromium, which is the element most important for formation of
the surface oxide layer that provides them their outstanding corrosion resistance. They also contain 12–15 wt.% of nickel and
2–3 wt.% of molybdenum [26]. In combination with Ni, Cr forms a complex, tenacious oxide barrier ensuring not only excel-
lent corrosion protection, but also stability of the austenitic phase. Furthermore, thin adhesive oxide layers on surfaces of
implanted components provide highly compatible surfaces for protein absorption and cellular interaction which support
healing of damaged tissues. On the other hand, chromium tends to form precipitates in the presence of carbon. Although
chromium-rich precipitates increase strength, their formation reduces the chromium content in their vicinity. This unfavor-
able effect can be reduced either with additions of molybdenum, which has a stronger affinity to carbon and tends to form
carbides more readily than Cr, or by minimizing the carbon content, which also prevents accumulation of carbides at grain
boundaries [20]. For these reasons, carbon content in such steels is usually kept below 0.03 wt.%. A typical example of a
microstructure of cast and solution treated 316LN stainless steel with low carbon content and high Cr and Ni contents fea-
turing austenitic grains and numerous twins is shown in Fig. 2a [50]. Not only does Ni contribute to the formation of a pro-
tective oxide layer, but it also stabilizes the non-magnetic FCC austenitic structure and contributes to solid solution
strengthening. Austenite-stabilizing elements generally support formation of twins. Moreover, nickel increases the stacking
fault energy (SFE), which induces twin formation in higher numbers and with lower thicknesses. However, the great disad-
vantage of Ni is its harmfulness to the human body. This inspired the development of Ni-free stainless steels, though they are
slightly ferromagnetic [26]. From the viewpoint of austenite stabilization, Ni can partly be substituted by Mn, Co, Cu, C and
N; the basic chemical composition of Ni-free steels typically contains iron plus Cr, Mn, Mo and N.
Nitrogen is considered the most suitable substitute for nickel. It is a very strong austenite stabilizer and provides stainless
steels with the ability to be work hardened via very high strains without any occurrence of strain induced martensitic trans-
formation. Relatively high volumes of nitrogen also interstitially dissolve in the austenitic crystal lattice, which induces sig-
nificant solid solution strengthening [50]. In addition, it suppresses development of intermetallic phases, supports formation
of strengthening precipitates, and decreases the SFE (especially in combination with Mo since they create Mo-N atomic pairs
affecting the SFE value and dislocation substructure character). Nitrogen addition improves mechanical properties (i.e. fati-
gue strength and fracture toughness) and pitting corrosion resistance. Moreover, if the original passivation layer is damaged,
nitrogen supports re-passivation via creation of an oxy-nitride surface layer, which is highly stable especially in environ-
ments of chloride ions [26]. Nevertheless, elevated nitrogen content can result in precipitation of Cr2N chromium-rich
nitrides on original austenitic grain boundaries. Although they contribute to strengthening by restraining movement of dis-
locations and grain boundaries pinning, they are undesirable since their formation decreases Cr content in the surrounding
matrix and may cause brittleness by shifting the transition temperature. Therefore, the content of nitrogen is usually kept
below 0.9 wt.%.
Another typical Ni substitute providing similar structure and properties is Mn [51]. It also stabilizes the austenitic phase
and supports formation of twins by increasing the SFE. While the austenite-stabilizing effect is better for Ni, Mn increases the
SFE more substantially and also ensures a noticeable solid solution strengthening, although the ductility of steels having Ni
replaced by Mn is lower. An example of a 317L Ni-free stainless steel containing 15 wt.% Mn and 0.5 wt.% N is shown in
Fig. 2b [25]. When compared to the typical 316LN steel microstructure shown in Fig. 2a, no significant structural differences
caused by the elimination of nickel can be seen. Fig. 2 thus clearly depicts examples of the results of modern research to
develop highly biocompatible materials by replacing harmful elements with more biocompatible ones while preserving
desirable microstructures.

3.2. Cobalt-chromium alloys

Several basic types of CoCrMo alloys featuring slightly different chemical compositions have been designed specifically
for joint replacements [52]. The two types predominantly used for orthopedic implants are CoCrMo (ATSM F75 and F76)
and CoNiCrMo (ATSM F562) [20]. Cobalt, the basic element in all these alloys, can crystallize in two different lattices, low
temperature hexagonal close packed (HCP) e phase and high temperature FCC c phase (sometimes designated also as a
phase) [53]. The presence of two different lattices creates obstacles to the movement of dislocations and thus induces
strengthening. However, the ability of CoCrMo alloys to fully transform the crystal lattice to HCP is suppressed due to the
presence of significant amounts of alloying elements. Consequently, the alloys usually maintain a metastable FCC matrix
even at room temperature. Nevertheless, these alloys can also be cold processed to invoke FCC to HCP solid-state phase
transformation and additional strengthening from precipitation [54]. Compared to stainless steels, which also exhibit two
different crystal structures, both the lattices within Co-based alloys are close-packed and thus relatively dense. This provides
the alloys with enhanced mechanical properties, such as excellent resistance to fatigue [26].
Co-based alloys usually contain 26–30 wt.% of Cr, which favorably influences their biocompatibility and corrosion resis-
tance via a strong tendency to create a thin Cr2O3 surface oxide film [55,56]. The addition of 5–10 wt.% of Mo primarily sup-
ports corrosion resistance and grain size refinement, although it also enhances solid solution strengthening via development
L. Kunčická et al. / Progress in Materials Science 88 (2017) 232–280 239

of the Co3Mo intermetallic phase [43,48]. Tungsten favorably influences the distribution of carbides, as well as their size and
degree of segregation at grain boundaries. CoNiCrMo alloys also contain Ni, the content of which is usually between 25 and
37 wt.% [20]. Although Ni substantially improves the corrosion resistance, high contents are controversial due to the ten-
dency for Ni to cause an inflammatory response in humans. All the main alloying elements also contribute to strengthening
via solid solution and precipitation of carbides. Precipitation strengthening can also be supported by additions of nitrogen
stimulating formation of chromium-rich nitrides (CrN, Cr2N) [24]. However, nitrogen is typically introduced into the struc-
ture via a surface treatment and thus its increased content is often limited only to surface and undersurface areas of the trea-
ted components.
The alloy processing technology and subsequent surface treatments significantly influence the type of microstructure and
consequent mechanical properties. CoCrMo components are usually either cast, or wrought. Microstructures of cast alloys
typically consist of lamellae with mixtures of hard phases, while wrought alloys are usually heat treated and therefore typ-
ically feature precipitated carbides [54]. Some CoCr-based alloys, such as CoNiCrMo, can be aged, which leads to additional
precipitation of carbides and further improvement of mechanical properties. Researchers showed that the mechanical prop-
erties of high-carbon CoCrMo alloys are superior to low-carbon variants; for example, higher amounts of carbides favorably
influence wear characteristics [57,58]. Generally, fine carbides precipitated within grains increase strength and ductility,
while coarse carbides at grain boundaries increase brittleness and decrease ductility. These findings shifted the focus of met-
als research towards increasing the carbon content in these alloys, which resulted in the development of special high-carbon
variants of CoCrMo [52]. Their structures exhibit tendencies to form spherical carbides, which provide enhanced mechanical
properties. The Cr23C6 carbide supports work-hardening and consequently enhances wear resistance. On the other hand, for-
mation of Cr-rich and Mo-rich carbides, such as Cr21Mo2C6, decreases contents of the elements in the surrounding matrix.
This can result in a local decrease in corrosion resistance, although corrosion resistance depends strongly on the overall
chemical composition and therefore heat treatment induced carbide precipitation does not have to necessarily lead to its
reduction [59].
Suffice to say, owing to its enhanced biocompatibility and favorable mechanical properties resulting from the interaction
of two different lattices and notable contents of alloying elements, CoCr-based alloys are leading joint replacement bioma-
terials. Nevertheless, fabrication and finishing treatments have to be processed carefully in order to avoid unintentional
reduction of one property at the expense of another [20]. Most of the components are cast in their final shapes, typically
using lost wax casting, since work hardening imparted by additional deformation processing can lead to decreased fracture
toughness. However, CoCr-based components still exhibit the deleterious effects of wear and fretting corrosion, which may
cause release of heavy metal ions into physiological environments and loosening of the implant. For this reason, some alloys,
especially those with Ni, should be avoided in applications experiencing hard contacts and friction loading, where e.g. Ti-
based alloys can advantageously be used.

3.3. Titanium and its alloys

Similar to Co, Ti typically crystallizes in two different lattice structures, a low temperature HCP a phase and a high tem-
perature BCC b phase [17]. The allotropic transformation temperature is 882 °C, although it can be shifted by the influence
of processing conditions and additions of alloying elements with specific effects on phase compositions. According to their

Fig. 3. Microstructure of a Ti6Al4V forged-piece depicting the a (lighter) and a + b (darker) phase areas [60].
240 L. Kunčická et al. / Progress in Materials Science 88 (2017) 232–280

effect on lattice structures, the alloying elements can be denoted as a stabilizers (oxygen, nitrogen, carbon, aluminum, rare
earths, etc.) and b stabilizers (hydrogen, vanadium, molybdenum, niobium, iron, chromium, cobalt, etc.) [15].
Titanium alloys are commonly categorized by their phase content as a phase alloys, near a phase alloys, a + b phase alloys
and b phase alloys. Commercial purity titanium, CP Ti (ATSM F67), and Ti-based alloys with additions of a stabilizing ele-
ments maintain a single HCP a phase structure at room temperature. The chemical composition of these alloys can be quan-
tified by their aluminum equivalent (Aleq = Al + 1/3Sn + 1/6Zr + 10[O + C + 2 N]), which is kept lower than 9 wt.%. Near a
phase alloys usually contain up to 2 wt.% of b stabilizing elements and thus their structures typically contain small portions
of b phase. However, neither Ti-based a alloys, nor near a alloys, are used for joint implants due to their unsatisfactory
mechanical properties. CP Ti is used in bioengineering currently mostly used in dental applications or for fabrication of por-
ous coatings [20].
The structures and phase compositions of a + b alloys can vary depending upon the amounts of alloying elements and
influence of heat treatments. Microstructures and properties thus depend on the individual chemical compositions, temper-
atures of possible solution and aging treatments and subsequent cooling rates. All the a + b alloys typically contain fine par-
ticles of both the a and b phases. The effect of b stabilizing elements on the stability of b phase can be quantified by the
molybdenum equivalent (Moeq = Mo + 0.6V + 0.44 W + 028Nb + 0.22Ta + 1.25Cr + 1.25Ni + 1.7Co + 2.5Fe). Generally, a higher
Mo equivalent value causes a decreased b to a transition temperature. The alloys usually also contain a stabilizing elements
dissolved interstitially in solid solution [17]. The element providing the highest interstitial strengthening effect is nitrogen;
the effects of carbon and oxygen are less intensive.
The most widely used bio-applicable Ti-based a + b alloy is Ti6Al4V (ASTM F136), the properties of which can vary sig-
nificantly due to the presence of both, an a stabilizer (Al) and a b stabilizer (V) [17,20]. An example of a typical forged
Ti6Al4V microstructure containing light a grains and darker a + b regions is depicted in Fig. 3 [60]. The presence of a phase
imparts good strength and resistance to oxidation, whereas b phase ensures better ductility of the final microstructure. On
the other hand, the hard and less ductile a phase increases the elastic modulus. Despite the fact that mechanical properties
of a + b alloys are generally closer to the properties of bone when compared to the afore mentioned alloys, they can still be
improved [15]. Thus, the research and development in the field of Ti-based alloys is at present directed at two main goals.
The first goal is to alter the chemical compositions of the alloys to replace controversial elements. This trend has already
resulted in the development of a + b alloys having the harmful aluminum and/or vanadium replaced by other elements
(e.g. Ti6Al7Nb and Ti5Al2.5Fe). The second goal is to develop alloys with properties as close to those of bone as possible
(to reduce differences between elastic moduli of bone and implanted materials). These efforts led to the development of
Ti-based alloys with structures consisting completely of b phase.
Ti-based b phase alloys maintain this phase even at room temperature, but their microstructures can also contain small
portions of a phase. However, the b phase structure is metastable and can be partially transformed to a phase or martensite
under certain conditions. These alloys are primarily alloyed with Nb, Zr, Mo and Ta, although other strong b stabilizers can
also be added in order to increase certain mechanical properties (e.g. hardenability can be improved by Fe and Cr additions)
[15]. However, chemical compositions containing elements with high melting points bring about processing problems, such
as difficult melting and solidification [17]. Moreover, casting must be performed under a protective atmosphere or vacuum
due to the very high affinity of titanium to oxygen. Such issues led to the introduction of powder metallurgy (PM) processes
into the production of b phase alloys. PM eliminates the need for high processing temperatures, as well as for protective
atmospheres. On the other hand, the affinity of Ti to oxygen can be used to improve certain properties. Due to the high

Fig. 4. Detail of the structure of Trabecular MetalTM [31].


L. Kunčická et al. / Progress in Materials Science 88 (2017) 232–280 241

oxygen solubility, surfaces of Ti components can be easily hardened via inward diffusion of oxygen during heating in air,
which also results in generation of TiO2 surface passivation layer, increasing the biocompatibility and corrosion resistance
of implanted components [17]. However, the TiO2 layer consists of a-phase, which deteriorates ductility and fatigue
strength, although the latter can be increased by additions of Y2O3 restricting dislocation movement and thus causing dis-
persion strengthening [61]. Nevertheless, Ti-based alloys have generally the lowest wear resistance among the above-
mentioned alloys and their application remains limited to components avoiding hard contacts.
The strength of b phase alloys is generally lower than of a and a + b alloys, however, it can be improved via precipitation
of carbides induced by thermo-mechanical treatments. On the other hand, precipitation increases the elastic modulus,
although this can be compensated by additions of other b stabilizing elements, such as tin [62]. Alloys featuring b structure
have generally lower moduli and are also more ductile than other Ti-based alloys due to the fact that most b stabilizing ele-
ments do not create intermetallic structures, but eutectoid phases [15]. The main aim in the development of b phase alloys is
to find chemical compositions that ensure the lowest possible elastic modulus and also good biocompatibility. Various
recent experimental investigations have evaluated several binary (TiMo [63,64], TiTa [65,66]), ternary (TiMoNb [67], TiMoSi
[68], TiNbSn [69], TiNbTa [70]), as well as quaternary (TiNbTaZr [71,72], TiZrTaNb [73], TiNbMoSn [74], TiNbZrSn [75])
alloys. Modern alloys should also be suitable for load bearing permanent implants.
One of the most notable findings of contemporary research is the influence of Ta on the elastic modulus of binary TiTa
alloys, which varies with Ta content [65,76]. Ta generally increases biocompatibility and corrosion resistance of implants,
which favor its usage without alloying with titanium.

3.4. Tantalum alloys

Considering limitations of all the above mentioned alloys (high elasticity modulus, low volumetric porosity, low frictional
characteristics, high wear rate, etc.), tantalum is one of the most promising emerging bio-applicable materials. Ta is able to
crystallize in two different lattice structures, BCC a phase and tetragonal metastable b phase [77]. The transformation tem-
perature is variable within a wide range and depends on chemical composition and presence of impurities, especially on the
content of oxygen [20,78].
Its high affinity to oxygen provides Ta with the ability to develop a thick (5 lm), stable, protective Ta2O5 surface layer
with excellent corrosion resistance [77]. Owing to the superb corrosion resistance, good biocompatibility and relatively inert
in vivo behavior, bulk Ta-based materials with additions of carbon have recently started to be used for orthopedic purposes.
They are already used for suture wires, plates, and spinal implants [79,80]. Nevertheless, despite its excellent behavior and
superior biological properties, such as proliferation and differentiation of human osteoblasts, Ta is not considered to be suit-
able for fabrication of large implants primarily due to its high price, high density, and a relatively high elasticity modulus
[81]. To overcome these disadvantages, porous forms of Ta have been developed [82]. They exhibit sufficient mechanical
properties, low elastic modulus and high resistance to frictional wear. On the other hand, the strength of porous structures
is low, which predisposes them only for non-load-bearing applications and coatings.
Ta coatings increase the biocompatibility and wear resistance of components fabricated from other bio-applicable alloys.
They are advantageous especially for CoCr-based components experiencing hard contacts, since coatings reduce wear and
the consequent release of heavy metal ions. Porous Ta materials are typically fabricated using carbon skeletons with high
porosities (98%) consisting of repeating dodecahedrons, on which CP Ta is deposited using chemical vapor deposition/
infiltration technologies [31]. Probably the most successful porous Ta-based structure is Trabecular MetalTM, developed by
Zimmer Inc. (Zimmer, Trabecular Metal Technology, Inc., Parsippany, NJ). Its typical structure is depicted in detail in
Fig. 4. Carbon skeletons, similar in structure to cancellous bones, significantly contribute to elasticity modulus reduction.
Their application can be performed e.g. by a patented treatment of components in molten salts containing Ta halides, such
as K2TaF7 or other Ta chlorides, fluorides, and bromides [83]. The thickness of Ta layer on the skeleton is usually between 40
and 60 lm and the overall porosity and pore size usually vary between 75–85%, and 400–600 lm, respectively. The superior
corrosion resistance, high friction coefficient, high porosity and good mechanical properties predispose porous Ta materials
for use in challenging orthopedic applications, such as joints and bone components exposed to high loads, and components
requiring bone ingrowth [84]. At present, porous Ta structures are used to fabricate acetabular components for primary and
revision total hip arthroplasties (THA), tibial components, and patella buttons for total knee arthroplasties (TKA).

3.5. Zirconium alloys

Like Ta, Zr is among the most promising modern bio-applicable materials due to its high strength, fracture toughness and
resistance to corrosion, good biocompatibility and non-toxicity. Its overall biological behavior is similar to the behavior of Ti
and it is a component of several newly developed b phase Ti materials, such as TiNbTaZr [71,72], TiZrTaNb [73] and TiNbZrSn
[75]. Due to its ability to develop a surface bone-like apatite layer when inside a living organism, Zr is also a suitable bioac-
tive material. Research on Zr-based materials with minor additions of other elements has begun only recently.
Zr crystalizes in two lattice structures, a low temperature HCP a phase and a high temperature BCC b phase. The transi-
tion temperature is around 863 °C, although it can be modified by additions of various alloying elements [85,86]. b phase Zr
has a low elastic modulus, which motivates researchers to develop materials maintaining this high-temperature phase also
at room temperature. This can be achieved by additions of b stabilizing elements, such as Nb, Mo, Ti and Si [87]. Similar to Si,
242 L. Kunčická et al. / Progress in Materials Science 88 (2017) 232–280

Table 1
Orthopedic device alloys and their properties.

Material Processing E (GPa) YS (MPa) US (MPa) Elong. (%) RA (%) HV30 (–) FS (MPa)
SS
AISI 316L A1 2007 314T 588T 49 84 155 40010
A1 + CD1 – 922T – – – –
SS
Rex734 A1 – 584T 898T 39 52 289 –
A1 + CF1 – 1325T – – – – –
SS
Panacea558 A1 – 600T 923T 54 74 367
A1 + CF1 – 1419T – – – – –
SS
Biodur 108 HF2 – 586T 931T 52 75 – –
SS
X13CrMnMoN A5 – 590T 1030T 70 75 259 –
18-14-3
SS
24Cr-1N HF + CF3 – – 1032T 26 – – –
SS
24Cr-2Mo-1N HF + CF4 – – 1167T 45 – – –
SS
BIOSSN4 HF + CR6 – 559T 938T 54 64 248 –
F75 (Co-Cr-Mo) A9 210 470T 770T 15 – – 280
HIP9 250 840T 1280T – – – 800
Co31Cr4Mo HIP8 170 – 965ST – – 427 48010
Co-28Cr-6Mo CD17 210 470T 780T 8 – – –
F90 (Co-Cr-W-Ni) A9 210 550T 1100T – – – –
CD9 210 1610T 1900T – – – 590
F799 (Co-Cr-Mo) HF9 210 1000T 1500T 28 – – 750
F562 (Co-Ni-Cr-Mo) HF9 230 980T 1210T – – – 500
CD9 230 1500T 1800T 8 – – 740
A9 230 280T 600T 50 – – –
F563 (Co-Ni-Cr-Mo) CD9 230 980T 1180T – – – –
CD + A11 54 – 591B – – – 30011
Zr1Nb A12 58 604T 786T 6.5 – – –
Zr3Nb A12 75 750T 862T 2 – – –

Zr12Nb A12 86 644T 687T 23.6 – – –


Zr20Nb as cast13 29 857C 1046C – – – –
Zr20Nb7Ti as cast13 29 910C 1145C – – – –
Zr20Nb15Ti as cast13 30 1071C 1307C – – – –
Zr12Mo as cast14 33 1200C 1542C – – – –
Zr12Mo3Ti as cast14 33 1211C 1582C – – – –
Ta porous15 1.22 12.7C 21.8C – – – –
Ta CD16 – 300T 316T 41 – – 365
Ta A16 – 234T 283T 53 – – 335
Zr8.8Si0.9Nb as cast18 29 850C 1189C 8.6 – – –

SS – stainless steel.
A – annealing.
CD – cold deformation.
CF – cold forging.
HF – Hot forging.
HIP – hot isostatic pressing.
CR – cold rolling.
US - Ultimate Strength.
RA – reduction of area.
T - in tension.
C - in compression.
S – In shear.
B - in bending.
E – Young’s modulus.
FS - fatigue strength.
1
U. Thomann and P. Uggowitzer, Wear–corrosion behavior of biocompatible austenitic stainless steels, Wear 239, 2000, 48–58.
2
C. Kraft, B. Burian, L. Perlick, M. Wimmer, T. Wallny, O. Schmitt and O. Diedrich, Impact of a nickel-reduced stainless steel implant on striated muscle
microcirculation: A comparative in vivo study, J Biomed Mater Res, 57, 2001, 404–412.
3
D. Kuroda, T. Hanawa, T. Hibaru, S. Kuroda and M. Kobayashi, Mechanical Properties and Microstructures of a Thin Plate of Nickel-Free Stainless Steel with
Nitrogen Absorption Treatment, Mater Trans 44, 2003, 1363–1369.
4
A. Yamamoto, Y. Kohyama, D. Kuroda and T. Hanawa, Cytocompatibility evaluation of Ni-free stainless steel manufactured by nitrogen adsorption
treatment, Mater Sci Eng C 24, 2004, 737..
5
S. Koch, R. Buscher, I. Tikhovski, H. Brauer, A. Runiewicy, W. Dudyinski and A. Fischer, Mechanical, Chemical and Tribological Properties of the Nickel-free
High-Nitrogen Steel X13CrMnMoN18-14-3, Materialwiss Werkst 33, 2002, 705–715.
6
R. Yibin, Y. Ke and B. Zhang, In vitro study of platelet adhesion on medical nickel-free stainless steel surface, Mater Lett 59, 2005, 1785–1789.
7
Q. Chen and G.A. Thouas, Metallic implant biomaterials, Mater Sci Eng R 87 , 2015, 1–57. doi:10.1016/j.mser.2014.10.001.
8
B. Henriques, M. Gasik, J.C.M. Souza, R.M. Nascimento, D. Soares and F.S. Silva, Mechanical and thermal properties of hot pressed CoCrMo-porcelain
composites developed for prosthetic dentistry, J Mech Behav Biomed Mater 30, 2014, 103–110. doi:10.1016/j.jmbbm.2013.10.023.
9
J.R. Davies (Ed.), Handbook of Materials for Medical Devices, ASM International, Materials Park, Ohio, 2003, pp. 21–50.
10
S. Teoh, Fatigue of biomaterials: a review, Int J Fatigue 22, 2000, 825–837. doi:10.1016/S0142-1123(00)00052-9.
11
S.A. Nikulin, V.A. Markelov, A.Y. Gusev, T.A. Nechaykina, A.B. Rozhnov, S.O. Rogachev and M. Yu. Zadorozhnyy, Low-cycle fatigue tests of zirconium alloys
using a dynamic mechanical analyzer, Int J Fatigue 48, 2013, 187–191. doi:10.1016/j.ijfatigue.2012.10.019.
12
R. Kondo, N. Nomura, Suyalatu, Y. Tsutsumi, H. Doi and T. Hanawa, Microstructure and mechanical properties of as-cast Zr-Nb alloys. Acta Biomater 7,
L. Kunčická et al. / Progress in Materials Science 88 (2017) 232–280 243

2011, 4278–4284. doi:10.1016/j.actbio.2011.07.020.


13
L. Nie, Y. Zhan, T. Hu, X. Chen and C. Wang, b-Type Zr-Nb-Ti biomedical materials with high plasticity and low modulus for hard tissue replacements, J
Mech Behav Biomed Mater 29, 2014, 1–6. doi:10.1016/j.jmbbm.2013.08.019.
14
L. Nie, Y. Zhan, H. Liu and C. Tang. Novel b-type Zr–Mo–Ti alloys for biological hard tissue replacements, Mater Des 53, 2014, 8–12. doi:10.1016/
j.matdes.2013.07.008.
15
R. Wauthle, J. Van der Stok, S. Amin Yavari, J. Van Humbeeck, J.-P. Kruth, A.A. Zadpoor, H. Weinans, M. Mulier and J. Schrooten, Additively manufactured
porous tantalum implants, Acta Biomater 14, 2014, 217–225. doi:10.1016/j.actbio.2014.12.003.
16
M. Papakyriacou, H. Mayer, H. Plenk and S. Stanzl-Tschegg, Cyclic plastic deformation of tantalum and niobium at very high numbers of cycles, Mater Sci
Eng A 325, 2002, 520–524. doi:10.1016/S0921-5093(01)01446-0.
17
B.R. Levine, S. Sporer, R.A. Poggie, C.J. Della Valle and J.J. Jacobs, Experimental and clinical performance of porous tantalum in orthopedic surgery,
Biomaterials 27, 2006, 4671–4681. doi:10.1016/j.biomaterials.2006.04.041.
18
C. Li, Y. Zhan and W. Jiang, Zr–Si biomaterials with high strength and low elastic modulus, Mater Des 32, 2011, 4598–4602. doi:10.1016/
j.matdes.2011.03.072.

Nb strongly affects transformation of the a phase to a final b phase structure [85]. Zr with additions of Mo and Ti has been
shown to completely maintain a b phase structure even at room temperature [87]. Additions of Ti also cause grain refine-
ment, although the final grain size depends also on the cooling rate and method. Mo and Nb form solid solutions and thus
contribute to solid solution strengthening by restricting the movement of dislocations; additions of Mo also strengthen the
material via introduction of lattice distortions.
Zr is also able to form bulk amorphous structures and metallic glasses, which typically have high strength and fatigue
resistance, low elastic modulus and good wear and corrosion resistances [88]. However, amorphous structures usually con-
tain controversial elements, such as Al, Cu, Co and Ni. For example, ZrAlCo glasses are promising despite their controversial
chemical composition, since they can easily be passivated, which lowers the possibility of releasing metal ions into tissue.
Nevertheless, glasses can also be alloyed with elements with sufficient biocompatibility, such as Nb and Ag, the additions
of which improve the glass forming ability and mechanical properties [88–90].
Suffice to say, achieving/maintaining the required microstructures of bio-applicable materials is essential, since they
impart to alloys their characteristic mechanical and biological properties.

4. Mechanical properties

With every movement, the joints of a human body are exposed to external influences such as loading forces and friction.
Spine members and lower extremity joints are probably subjected to the most challenging conditions [15]. The force acting
on the hip joint during a step (lower extremity is deflected) is several times one’s body weight [91], while the knee joint
experiences compressive loads as high as 4–8 times body weight during casual daily activities [92,93]. Long-time exposures
of joints to inappropriately high or disproportional loads are responsible for most of the degenerative processes, which can
finally result in the need to replace an implant.
In general, artificial joints are anchored to bone tissue. Therefore, the properties of joint replacement materials should
match the properties of human bones as closely as possible. Metallic implants that directly integrate with bone usually pos-
sess sufficient strengths, but their elastic moduli are typically too high. This results in the stress shielding effect, causing
aseptic loosening of the implants during long-term applications [15]. However, loosening of an implant can also be a con-
sequence of its low osseointegration ability or low wear resistance. On the other hand, low biocompatibility and low corro-
sion resistance can lead to inflammations or fibrous encapsulations in surrounding tissues [3]. Dental and skeletal bone
implants also suffer from cyclic loading, which can result in fatigue failure [93].
Mutual interactions of the individual materials within a single surgical replacement are important as well. Although the
basic designs of a total hip joint replacement are similar to designs developed in the 1970s, the selection of available mate-
rials has become wider [20]. For example, a modern THA design contains Ti and CoCr materials, as well as UHMWPE and
ceramics, while knee joint replacements typically are made from stainless steel, CoCrMo and Ti-based alloys supplemented
with polymers (UHMWPE, PMMA and others) [15,46,94]. For these reasons, the following sub-sections focus on the most
important properties of human joint replacement materials. For a better clarity, the important properties of the most com-
monly used joint replacement materials are summarized in Table 1, with the exception of Ti-based alloys, the properties of
which are further summarized in Table 2 (Section 4.2).

4.1. Strength

Mechanical strength enables bone tissue to withstand external forces and physiological loadings. Although the individual
values vary according to the bone type, the compressive yield strength (YS) typically ranges from 130 to 240 MPa and the
ultimate tensile strength (UTS) can be up to 280 MPa [15,95,96]. For example the strength of long bones (leg, arm) varies
between 100 and 200 MPa, while for vertebrae it ranges between 1 and 10 MPa [15]. To ensure long-time durability of an
implant, its strength should be like that of the bone. Excessive strength can cause deterioration of the surrounding bone.
On the other hand, an implant with poor strength will not be able to endure all the in vivo stresses [97].
244 L. Kunčická et al. / Progress in Materials Science 88 (2017) 232–280

Table 2
Titanium alloys and their properties.

Alloy Processing Phase E (GPa) YS (MPa) UTS (MPa) FS (MPa) BS (MPa) BM (GPa)
cpTi com. sup. a 105 485 550 – 8006 1006
cpTi annealed2 a – 480 550 350 – –
cpTi cast14 a 105 530 640 325 – –
6
Ti10Sn – – – – 12806 1306
Ti12Cr13 sol.treat 68 – 78027 – – –
Ti12Cr13 cold rolled 86 – 93027 – – –
Ti7.5Mo14 cast 80 665 997 300 – –
Ti10Mo19 sintered* 30 – 254 – – –
Ti15Mo18 sintered* 32 – 310 – – –
Ti6Mo4Sn15 drawn 50 785 825 – – –
Ti6Mo4Sn15 annealed 39.5 514 875 – – –
Ti10Mo7Nb17 cast 28.4 1854 1918 – 701 –
Ti10Mo7Nb17 cast 27.3 1685 1808 – 818 –
Ti10Mo7Nb17 cast 24.7 1404 1717 – 1489 –
NiTi12 48 – – – – –
Ti6Al4V com. sup. a+b 110 800 860 – – –
Ti6Al4V cast14 a+b 117 850 1137 360 – –
Ti6Al4V annealed1 a+b – 680 780 400 – –
Ti6Al4V hot forged a+b 11012 900 1000 600 – –
Ti6Al7Nb com. sup. a+b 105 – 98012 – – –
Ti6Al7Nb annealed3 a+b – 800 900 500 – –
Ti5Al2.5Fe com. sup. a+b 110 – – – – –
Ti5Al2.5Fe cast3 – 820 900 425 – –
Ti5Al2.5Fe annealed3 – 780 860 725 – –
Ti7.5Nb4Mo20 hot forged b+a+x 120 520 1010 – – –
Ti7.5Nb4Mo2Sn20 hot forged b+a+x 87 700 950 – – –
Ti13Nb13Zr com. sup. b metastable 79 – 99012 – – –
Ti13Nb13Zr14 cast b 77 656 945 275 – –
Ti13Nb13Zr annealed2 b – 900 1030 500 – –
Ti16Nb13Ta4Mo12 b 91 – – – – –
Ti24Nb4Zr7.9Sn9 b 42 – – – – –
Ti25Nb22 b – – – – 1100 90
Ti25Nb11Sn22 b 45.623 – – – 1780 53
Ti40Nb2Sn quenched b 55 – – – – –
Ti29Nb13Ta4.6Zr13 sol. treat b 61 – – – – –
Ti29Nb13Ta4.6Zr13 cold rolled b 61 – – – – –
Ti35Nb4Sn annealed b 55 380 450 – – –
Ti35Nb4Sn21 cold rolled b 42 600 660 – – –
Ti35Nb7.9Sn24 annealed b 62 420 440 – – –
Ti35Nb7.9Sn24 cold rolled b 58 580 605 – – –
Ti35Nb4Sn6Mo9Zr25 hot forged b 68 750 780 – – –
Ti23Nb0.7Ta2Zr0.5N cold rolled b 50 500 660 – – –
Ti11.5Mo6Zr4.5Sn com. sup. b 79 – – – – –
Ti11.5Mo6Zr4.5Sn annealed3 b – 620 690 525 – –
Ti12Mo6Zr2Fe12 com. sup. b 80 – 1080 – – –
Ti15Mo5Zr3Al com. sup. b 80 – – – – –
Ti15Mo5Zr3Al annealed b 8012 900 930 540 – –
Ti15Mo3Nb0.3O com. sup. b 82 – – – – –
Ti15Mo3Nb0.3O annealed4 b – 1020 1020 490 – –
Ti10Mo1.25Si4Zr16 cast b 23 825 1412 – – –
Ti10Mo1.25Si13Zr16 cast b 32 1024 1417 – – –
Ti35Nb10Zr cold swaged11 b 50 850 900 – – –
Ti35Nb10Zr aged11 b+a 90 1230 1330 – – –
Ti35Nb2.5Ta7.5Zr cold swaged11 b 60 710 790 – – –
Ti35Nb2.5Ta7.5Zr aged11 b+x 80 970 1020 – – –
Ti35Nb5Ta5Zr cold swaged11 b 63 770 790 – – –
Ti35Nb5Ta5Zr aged11 b+x 80 960 1000 – – –
Ti35Nb7.5Ta2.5Zr cold swaged11 b 59 740 785 – – –
Ti35Nb7.5Ta2.5Zr aged11 b+x 79 950 990 – – –
Ti35Nb10Ta cold swaged11 b 58 780 810 – – –
Ti35Nb10Ta aged11 b+a+x 78 990 1020 – – –
Ti35Nb5Ta7Zr com. sup. b 55 – – – – –
Ti35Nb5Ta7Zr annealed4 b – 530 590 265 – –
Ti35Nb5Ta7Zr cold rolling7 b 55 – 800 – – –
Ti35Nb5Ta7Zr HPT 5 turns7 b 61 – 1100 – – –
Ti35Nb5Ta7Zr0.4O com. sup. b – – – – – –
Ti35Nb5Ta7Zr0.4O annealed4 b 66 976 1010 450 – –
Ti29Nb13Ta4.5Zr – b 65 – – – – –
Ti35Nb5Ta7Zr0.4O annealed5 b – 400 420 325 – –
L. Kunčická et al. / Progress in Materials Science 88 (2017) 232–280 245

Com.sup. – commercially supplied.


FS - fatigue strength.
BM - bending modulus.
BS-bending strength.
1
M. Semlitsch and H.G. Willert, Properties of implant alloys for artificial hip joints, Medical and Biological Engineering and Computing 18, 1980, 511–520.
2
J.C.M. Li, Microsctructure and properties of materials, World Scientific 2, 2000, 49–55.
3
B. Boyer, G. Welsch, and E.W. Collings, Materials Properties Handbook: Titanium Alloys, fourth ed., ASM International, 2000, pp. 790–810.
4
R. Narayan, Materials for medical devices, fundamentals of medical implant materials, ASM Handbook 23, 2012, pp. 6–16.
5
M. Niinomi and M. Nakai, Titanium-based biomaterials for preventing stress shielding between implant devices and bone, International Journal of Bio-
materials, 2011, 836587–10.
6
H. Hsueh-Chuan, W. Shih-Ching, H. Yu-Sheng, and H. Wen-Fu, Mechanical properties and deformation behavior of as-cast Ti–Sn alloys, J Alloy Compd 479,
2009, 390–394.
7
H. Yilmazer, M. Niinomi, M. Nakai, K. Cho, J. Hieda, Y. Todaka and T. Miyazaki, Mechanical properties of a medical b-type titanium alloy with specific
microstructural evolution through high-pressure torsion, Mater Sci Eng C 33, 2013, 2499-2507.
8
H. Matsumoto, S. Watanabe and S. Hanada, Strengthening of low Young’s modulus beta Ti–Nb–Sn alloys by thermomechanical processing. In: Proc.
Materials & Processes for Medical Devices Conference, ASM International, 2006, pp. 9–14.
9
Z. Guo, J. Fu, Y.Q. Zhang, Y.Y. Hu, Z.G. Wu, L. Shi, M. Sha, S.J. Li, Y.L. Hao and R. Yang, Early effect of Ti–24Nb–4Zr–7.9Sn intramedullary nails on fractured
bone, Mater Sci Eng C 29, 2009, 963-968.
10
M. Nakai, M. Niinomi, X.F. Zhao and X.L. Zhao, Self-adjustment of Young’s modulus in biomedical titanium alloy during orthopaedic operation. Mater Lett
65, 2011, 688–690.
11
J. Málek, F. Hnilica, J. Veselý, B. Smola, S. Bartáková and J. Vaněk, The influence of chemical composition and thermo-mechanical treatment on Ti–Nb–Ta–
Zr alloys, Mater and Des 35, 2012, 731–740.
12
M. Geetha, A. K. Singh, R. Asokamani, and A.K. Gogia, Ti based biomaterials, the ultimate choice for orthopaedic, implants – A review, Prog Mater Sci 54,
2009, 397-425.
13
M. Niinomi, M. Nakai and J.Hieda, Development of new metallic alloys for biomedical applications, Acta Biomater 8, 2012, 3888–3903.
14
Ch.W. Lin, Ch.-P. Ju and J.H.Ch. Lin, A comparison of the fatigue behavior of cast Ti–7.5Mo with c.p.titanium, Ti–6Al–4V and Ti–13Nb–13Zr alloys,
Biomaterials 26, 2005, 2899–2907.
15
Y. Sutou, K. Yamauchi, T. Takagi, T. Maeshima and M. Nishida, Mechanical properties of Ti–6 at.% Mo–4 at.% Sn alloy wires and their application to medical
guidewire, Mater Sci Eng A 438–440, 2006, 1097–1100.
16
Y. Zhan, Ch. Li and W. Jiang, b-type Ti-10Mo-1.25Si-xZr biomaterials for applications in hard tissue replacements, Mater Sci Eng C 32, 2012, 1664–1668.
17
L.J. Xu, Y.Y. Chen, Zh.G. Liu and F.T. Kong, The microstructure and properties of Ti–Mo–Nb alloys for biomedical application, J Alloy Compd 453, 2008, 320–
324.
18
Y.H. Li, R. B Chen, G. Qi, Z. T. Wang and Z.Y. Deng, Powder sintering of porous Ti–15Mo alloy from TiH2 and Mo powders, J Alloy Compd 485, 2009, 215–218.
19
Z. Gao, Q. Li, F. He, Y. Huang and Y. Wan, Mechanical modulation and bioactive surface modification of porous Ti–10Mo, alloy for bone implants, Mater Des
42, 2012, 13–20.
20
D.C. Zhang, S. Yang, M. Wei, Y.F. Mao, C.G. Tan and J.G. Lin, Effect of Sn addition on the microstructure and superelasticity in Ti–Nb–Mo–Sn Alloys, J Mech
Behav Biomed 13, 2012, 156–165.
21
T.K. Jung, H. Matsumoto, T. Abumiya, N. Masahashi, M.S. Kim and S. Hanada, Mater Sci Forum 631–632, 2010, 205–210.
22
H. Hsueh-Chuan, W. Shih-Ching, H. Shih-Kuang, S. Jhen-Yi and H. Wen-FuHo, The structure and mechanical properties of as-cast Ti–25Nb–xSn alloys for
biomedical applications, Mater Sci Eng A 568 , 2013, 1–7.
23
K. Miura, N. Yamada, S. Hanada, T.K. Jung and E. Itoi, The bone tissue compatibility of a new Ti-Nb-Sn alloy with a low Young’s modulus, Acta Biomater 7,
2011, 2320–2326.
24
H. Matusmoto, S. Watanabe and S. Hanada, Beta TiNbSn alloys with low Young’s modulus and high strength, Mater Trans 46, 2005, 1070–1078.
25
S.-J. Dai, Y. Wang, F. Chen, X.-Q. Yu and Y.-F. Zhang, Influence of Zr content on microstructure and mechanical properties of implant
Ti35Nb4Sn6MoxZr alloys, T Nonferr Metal Soc 23, 2013, 1299-1303.
26
I. Raluca, G. Doina-Margareta, M. Valentina, P. Osiceanu, S. Dinescu, T. Gloriant and A. Cimpeana, In vitro bio-functional performances of the novel
superelastic beta-type Ti–23Nb–0.7Ta–2Zr–0.5N alloy, Mater Sci Eng C 35, 2014, 411–419.
27
M. Nakai, M. Niinomi, X. F Zhao, and X. L. Zhao. Self-adjustment of Young’s modulus in biomedical titanium alloy during orthopaedic operation. Mater Lett
65, 2011, 688–690.
*
85% density.

The yield strength of 316L stainless steels is 210–280 MPa, while the UTS usually ranges between 540 and 1000 MPa
[15,26]. The strength is higher for Ni-free steels, for which both the YS and UTS increase linearly with increasing content
of interstitially dissolved N at room temperature and constant grain size [26]. The new Ni-free P558 steel features a very high
strength together with a high work hardening capacity, while the recently developed BioDur 108 steel with the YS of
600 MPa and UTS exceeding 900 MPa is considered as a promising alternative to 316L [9]. The strength of steels can gen-
erally be improved by cold deformation processing. For example, the UTS of BIOSSN4 steel, which is 2–3 times higher than of
316L can be increased up to 1500 MPa by severe cold processing [98].
Most of the CoCr-based alloys have the YS exceeding 500 MPa and UTS ranging between 900 and 1540 MPa, which is gen-
erally comparable to the properties of newly developed steels. The strength is generally higher for wrought than for cast
work pieces [15]. Cast CoCrMo used in biomedical applications (ASTM F75) possesses YS ranging from 450 to 520 MPa
and UTS levels between 650 and 890 MPa, while forged CoCrMo (ASTM F799) has YS levels between 900 and 1030 MPa
and UTS levels from 1400 to 1590 MPa. The strengths of other CoCrMo types alloyed with elements such as W and Ni
(e.g. ASTM F90 and F562) are even higher, especially after thermo-mechanical processing. Particularly notable is the tensile
strength of cold processed and aged ASTM F562 CoNiCrMo alloys reaching almost 1800 MPa. It is the highest value among all
the materials used for joint implants, however, the highest specific strength among all the metal biomaterials can generally
be observed in Ti-based alloys.
246 L. Kunčická et al. / Progress in Materials Science 88 (2017) 232–280

Mechanical properties and strength of Ti-based alloys are strongly dependent on their chemical composition and phase
structure [9]. They are comparable to stainless steels, however, they are lower than for Co-based alloys. The YS of CP Ti
ranges typically from 170 to 480 MPa, while its UTS varies from 240 to 550 MPa depending on the impurity content [15].
For Ti6Al4V, the UTS can reach 930 MPa (or slightly less for similar vanadium-free alloys), while the YS is typically between
795 and 860 MPa [15]. b phase alloys generally have a UTS from 600 to 1100 MPa and their YS achieves similar maximum
values. The strength of Ti alloys increases with increasing oxygen content and can generally be enhanced via various treat-
ments, such as solution treatment with subsequent aging, thermo-mechanical processing and SPD technologies [9,99,100].
For example, Ti12Mo3Nb exhibited mean YS and UTS values of 449 MPa and 747 MPa, respectively, after vacuum casting and
solution treatment [101], while both the YS and UTS ranged between 600 and 800 MPa for cast, forged and solution treated
Ti35Nb6Mo4Sn. These levels gradually increased with increasing additions of Zr [102]. Comparison of b phase Ti26Nb and
Ti24Nb0.5O has shown the ability of Ti alloys to be oxygen hardened; the latter exhibited twice as high UTS as the former
(810 MPa as compared to 420 MPa) [103].
Similar to titanium, the strength of tantalum depends on its structure, processing technology and chemical composition.
Particularly important is the content of impurities (N2, O2, C), since even a small amount of interstitial atoms interacts with
dislocations and can cause changes in the stress-strain response during cyclic loading [104]. The standards for maximum
contents of admixtures in pure Ta for surgical purposes are given in ASTM F560 [15]. Considering the processing technology,
the strength is generally higher for cold-worked components, for which the YS reaches 345 MPa, while YS of annealed com-
ponents can be as low as 140 MPa [104]. Similarly, the UTS of annealed components is lower (approximately 205 MPa) than
of cold-worked ones (around 480 MPa). As shown, cold processing even increased the static strength of pure Ta to 813 MPa
[105]. Nevertheless, the strength is significantly different for porous Ta structures, such as Trabecular MetalTM, for which the
YS typically ranges between 35 and 51 MPa, while the UTS varies from 50 to 110 MPa [31].
As for zirconium, the strength also depends on the particular structure and chemical composition [86]. Pure Zr, as-cast
after arc melting, exhibits a YS of 304 MPa and UTS of 388 MPa. Nb addition usually results in a substantial increase in
strength, although the increase depends on the processing conditions and final treatment as well. In general, ZnNb b type
alloys used for modern joint implants have favorable mechanical properties. The YS of Zr with 3–24 wt.% Nb varies between
559 and 750 MPa, while the UTS ranges between 605 and 881 MPa. The strength can even be increased by additions of Ti and
Mo. As-cast Zr20Nb alloy has a YS of 857 MPa and compressive strength of 1046 MPa. The addition of 15 wt.% Ti increased
the values to 1071 MPa and 1307 MPa for YS and UCS, respectively [106]. The addition of 1 wt.% Mo into pure Zr resulted in a
UTS of 970 MPa [107]. Zirconium ceramics generally have relatively high strength. However, compression properties are
superior to tensile properties [20]. A typical UTS for ZrO2 in tension is 420 MPa, while in compression it reaches 7500 MPa.

4.2. Elastic stiffness

The elastic modulus of bones ranges between 5 and 30 GPa depending on bone type and direction of measurement, how-
ever, the modulus of metal implant materials is usually higher [95,96,108,109]. The difference between moduli of bone tis-
sue and implants causes the ‘‘stress shielding effect” – insufficient loading reduces stress on the bone, which causes
resorption and associated reduction of bone mass. The joint replacement then responds to external loads by micro-
motion and possibly by eventual loosening from the bone [110,111]. The modulus generally depends on structure, porosity,
and phase and chemical compositions, which can therefore be advantageously modified by alloying and thermo-mechanical
treatments [15,112].
The modulus of ASTM F138 316L stainless steel usually ranges from 190 to 210 GPa and can be decreased by nitrogen
additions. However, the content of nitrogen dissolved in the solid solution has to be sufficient (HNS steels), since a minor
addition results in precipitation of Cr-rich nitrides with almost negligible effect on the elastic properties [15,20,113]. The
addition of 2.3 wt.% N was shown to decrease the modulus of a slightly porous sample of an austenitic stainless steel under
165 GPa [114]. Greater porosity and processing by PM technologies can dramatically decrease the modulus. Bender et al.
[115] reported the modulus to be 85 GPa, 72 GPa and 62 GPa for samples of 302 stainless steel consolidated from powders
with the theoretical density of 79%, 74% and 70%, respectively.
The elastic modulus of pure cobalt is approximately 210 GPa in tension and 180 GPa in compression, which is similar to
nickel and iron [15]. The modulus of Co-based alloys usually ranges between 210 and 250 GPa and depends on the particular
microstructure especially for wrought components. For example, Liao et al. [54] reported the modulus of hard phases in cast
CoCrMo alloy to be 269 GPa, while the modulus of precipitated carbides in a wrought CoCrMo was 309 GPa. Moreover,
the precipitates were harder and had a lower plasticity. Even though the modulus of CoCr alloys is generally very high, it
can also be lowered by PM technologies. For example, the modulus of Co31Cr4Mo was 170 GPa after 10 min of hot compres-
sion [116].
As for Ti-based alloys, the modulus varies significantly with varying chemical and phase compositions. A summary of
hitherto developed low-modulus Ti-based alloys is given in Table 2. The modulus of commonly used Ti6Al4V is about
110 GPa, which is significantly lower than for steels and CoCr-based alloys, although still higher than for bones [69,117].
The modulus can further be lowered by replacing elements with high modulus (e.g. V) with elements with lower modulus
(Nb, Zr, Mo, Ta, etc.), which introduces b phase microstructure. The first b type alloy with low-modulus was Ti13Nb13Zr [9],
however, the alloy with probably the lowest modulus achieved so far is Ti35Nb (42 GPa) [118]. Nevertheless, one of the most
widely investigated alloys is the highly biocompatible TNTZ (Ti29Nb13Ta4.6Zr), for which the lowest reported modulus was
L. Kunčická et al. / Progress in Materials Science 88 (2017) 232–280 247

60 GPa (after solution heat treatment) [72]. A very low modulus has also been observed for TiNbSn alloys (40 GPa) [119].
Tin addition has generally a very favorable influence on the modulus. Elastic properties of binary TiV and TiNb alloys also
decrease with Sn additions. The value of 50.7 GPa was reported for Ti35Nb2.5Sn alloy [120], while a heat treated Ti-based
alloy with additions of Cu, Ni, Sn and Nb was reported to have a modulus of 54 GPa [117]. Other aspects that can influence
the elastic modulus are stability of the b phase and porosity. The value of 44 GPa reported for a Ti10Mo sample produced by
PM further decreased to 20 GPa after addition of 25% of NH4HCO3 [63]. A sintered Ti15Mo sample with a theoretical density
of almost 40% exhibited a modulus as low as 4.6 GPa [64]. Extremely low moduli of 0.3–2 GPa were observed for highly por-
ous Ti16Sn4Nb samples [121]. Finally, the modulus can also be influenced by SPD processing. For example, the high pressure
torsion (HPT) technology decreased the modulus of TNTZ from 64 GPa to 60 GPa [72], while accumulative roll bonding (ARB)
to 8 layers lowered the modulus of Ti10Zr5Nb5Ta from 58 GPa to 46 GPa [73]. Nevertheless, Ti6Al4V ELI is still the most
widely used Ti alloy in biomedicine, since most of the b phase alloys are still in the process of development and biological
testing [17].
As already mentioned, tantalum additions have different influences on the modulus of Ti alloys [65]. TiTa alloys with Ta
contents between 10 and 30 wt.% have a very low modulus, whereas Ta additions greater than 30 wt.% unfavorably increase
modulus and density of the alloys. The modulus reaches a value of 88 GPa for the composition of 50Ti50Ta and then
decreases to 67 GPa for 30Ti70Ta. With further additions of Ta, the modulus increases again and eventually reaches
186 GPa, the value of pure Ta. Probably the lowest modulus among TiTa alloys has been achieved for a rolled and solution
treated Ti25Ta (64 GPa) [122]; addition of 50 wt.% Ta led to its increase to 77–93 GPa [123].
Pure Ta is usually used in porous forms, which imparts a low modulus to the components [15]. Studies have shown the
modulus of porous Ta components fabricated using the LENSTM (Laser Engineered Net Shaping) process to range between 2
and 20 GPa for the open porosity between 55 and 27%, respectively [124]. The modulus of highly porous structures can be
lowered to 2.5–3.9 GPa [31], however, a modulus as low as 1.22 GPa has already been reported for a pure Ta component with
open porosity of 80% [125].
Notable is also the elastic modulus of Zr-based alloys, which is almost two times lower than for Ta [15]. However, it
depends on the phase composition of the alloy. While the modulus of a phase Zr reaches 100 GPa, for b phase Zr it is
60 GPa [15,87]. Silicon strongly influences the development of b phase and consequently the modulus; Zr8.8Si alloys con-
taining up to 1 wt.% of Nb have the modulus less than 30 GPa [85]. Likewise for the Ti-Ta system, the modulus of Zr-Nb alloys
decreases with increasing Nb content to approximately 74 GPa at 6 wt.% Nb and then increases to 90 GPa at 9 wt.% Nb [86].
The lowest value of 48.4 GPa features the composition of Zr20Nb, then it increases again with increasing Nb addition. A very
low modulus was also observed for Zr12Mo (33 GPa) [87].

4.3. Wear

Up to 80% of all the surgical revisions of joint prostheses must be performed due to loosening of the implant caused by
sliding wear. Three basic mechanisms of wear can be identified: adhesive wear (occurring when the atomic forces between
the surfaces of the materials subjected to load are stronger than their inherent properties), abrasive wear (resulting in abra-
sion of the softer material in the contact pair due to the friction of the harder one), and the third body wear mechanisms
(occurring when hard particles of an abraded brittle material get embedded in a soft surface) [15,126,127]. Wear mecha-
nisms and rates vary for congruent joints (hip, shoulder, etc.), where the stress distribution is more or less homogenous,
and incongruent joints (knee, ankle, etc.), where the stress distribution is inhomogeneous due to contacts of hard surfaces
[15]. Wear rate is usually the highest within a short time period after the implantation and continuously decreases until
reaching a more or less constant value [20,126].
Due to the relative motion of implanted components, wear debris inevitably generates at all sliding interfaces – between
polymers, ceramics, and metals – since all the materials have a certain finite roughness, as summarized in Table 3. Wear for
metal-on-metal and metal-on-ceramic designs is far lower than at interfaces with polymers; the largest amount of debris is
generated by polymers sliding against metals and ceramics. However, contact couples of two FCC metals are not favored due
to the possibility of occurrence of adhesion. The ceramic-on-ceramic designs exhibit the volumetric wear rate of approxi-
mately 0.04 mm3/year, while the rate for the best metal-on-metal design was 0.9 mm3/year [128]. Wear rates for
polymer-metal hip implant systems were reported to be between 4.5 mm3/year and 78 mm3/year [128–130].
Various types of contact couples produce debris particles that vary in size and shape. Debris particles provoke innate
inflammatory responses and potential rejection of the implant. Loose particles can moreover be transferred to surrounding

Table 3
Typical surface roughnesses of polymers, metals, and ceramics used in hip implants [2,130].

Material Type Material Average roughness Ra (lm)


Polymer Ultra High Molecular Weight Polyethylene 0.100–2.500
Metal Stainless Steel 0.010–0.050
CoCrMo
Ceramic Alumina 0.0008–0.015
Zirconia 0.001–0.005
248 L. Kunčická et al. / Progress in Materials Science 88 (2017) 232–280

tissue and stimulate harmful biological responses [20]. The presence of foreign particles invokes a complex cascade of cel-
lular and biochemical mechanisms, which are unique to the chemical and morphological characteristics of the debris. For
example, hard metal-on-metal designs usually produce small (submicron-size), round particles, while particles produced
by metal-on-polymer contact pairs are larger (micron-size) and elongated. The basics for characterization of particles pro-
duced in total joint simulation systems, abrasion testing machines, or harvested within clinical studies, are defined by stan-
dard methodologies (ASTM F1877).
The low hardness and weak passivation layer of stainless steels generally results in poor tribological properties and con-
sequently higher wear rates [131]. The modern Ni-free steels have generally higher wear resistance compared to 316L due to
their strain hardening ability [26,127]. Relative to a standard specimen of an A514 steel, the dry wear rate of the 316L steel
was almost 1, while the values for Rex734 and P558 Ni-free steels were slightly more than 0.8 and slightly less than 0.7,
respectively; P558 exhibited the best overall performance [132]. Pin-on-disc tests at the distance of 2 km showed the steady
state wear rate of 8.5  108 and 7  108 mm3 (25 and 30 lm/h) for CrNiMo and CrNiMnMoN steels, respectively, while a
CrMnMoN steel exhibited a wear rate of 2  109 mm3 [126].
Wear of steels can be decreased by PM technologies, additions of constituents with low wear rates, and applications of
coatings (surface treatment is further discussed in Section 5.1). The wear rate decreased with increasing HA content for a
PM–produced mixture of steel and hydroxyapatite (HA); from approximately 1500  106 mm3 for HA plus 60% steel pow-
der to approximately 320  106 mm3 for 100% HA (wear distance 1000 m) [127]. Of note is the change in the wear mech-
anism – abrasion to adhesion – with increasing steel powder content. A highly resistant CrB1.87 coating was shown to provide
Ortron 90 steel with a generally decreased wear rate [133]. Nevertheless, wear of steels is higher than for other biomaterials.
When compared to CoCrMo, 316L stainless steel has a 10 times higher wear rate and 1000 times larger debris particles [126].
Since high hardness correlates with high wear and scratch resistances, CoCrMo alloys (ASTM F75, F90, F562, F1537) are
favored for applications experiencing sliding wear due to among the highest hardness of any biomedical alloy (300–400 HV)
[134,135]. However, the wear resistance of CoCr-based alloys depends on their structure and the manufacturing process.
Wrought components generally wear more slowly than cast components and HCP structures wear less than FCC structures
[9,136]. For example, a forged CoCrMo component exhibited a wear rate of approximately 10  108 mg/Nm under the load
of 8.5 MPa [56]. Resistance to wear can be increased e.g. by development of a Cr2O3 surface layer, increasing nitrogen content
and precipitation of M23C6 carbides. However, various low-carbon CoCrMo alloys have exhibited very low wear rates as well
[52,126,137]. The abraded weight for an as-cast CoCrMoN sample decreased from 0.0073 g for 0 ppm of nitrogen to
0.0043 g for 2880 ppm of nitrogen [137]. Surface plasma nitriding can also decrease the wear rate of CoCrMo (see Sec-
tion 5.1). PM technologies can improve wear resistance [138]. A hot compacted CoCrMo sample exhibited a wear rate of
3.6  108, while for cast samples it was three times higher.
Ti-based alloys are generally not used for components articulating against each other due to their relatively low wear
resistance. This can be attributed primarily to the instability of their surface layer, low resistance to plastic shear and low
work hardening ability, however, harder martensitic structures exhibit better wear resistance. The wear rate for b type alloys
is generally lower than for harder a + b structures. The wear rate of the commonly used Ti6Al4V was reported to be
0.0004 mm3/m, slightly lower after water quenching [139]. Due to its high adhesiveness and low resistance to plastic shear,
TNTZ exhibited a higher volume loss when compared to a + b Ti64 during ball-on-disc tests (1.2  104 cycles at 60 RPM, load
4.9 N) [140]. b type Ti13Nb13Zr alloy exhibited a wear rate higher than Ti6Al4V during G77-83 ASTM standard testing
(0.33 mm3/m) [139]. On the other hand, some b phase alloys have wear rates lower than a + b alloys. When compared
to Ti6Al4V, a better wear resistance is exhibited for example by TiSiZr alloys and Ti-Ta alloys. For Ti6Si5Zr wear rates as
low as 0.58  104 mm3/Nm (ball-on-disc test at 160.8 m) have been documented [141]. Pin-on-disc tests (100,000 cycles,
load 4.9 N) showed the mass loss to be 0.03 g for Ti6Al4V, but only  0.02 g for Ti30Ta, 0.01 g for Ti10Ta and slightly more
than 0.005 g, for Ti70Ta [142]. Finally, the chemical composition and phase structure significantly influence the wear rate,
too. For example, resistance to wear by TNTZ increased with increasing Nb content (45  102 mm3 wear loss for 29 wt.%
Nb compared to 13  102 mm3 for 39 wt.% Nb, pin-on-disc tests with stainless steel at the distance of 30 km) [140].
Resistance to wear for Ti alloys can be enhanced using modern manufacturing technologies (PM) and surface modifica-
tions. The wear rate of as-cast CP Ti was reported to be 2  1012 m3/m, while Selective Laser Melting (SLM) decreased the
rate to 1.2  1012 m3/m [143]. The positive effect of thermal oxidation on improvement of wear resistance of Ti6Al4V and
CP Ti in corrosive environments has been shown by several studies [144–146]. Tests performed with Ti13Nb13Zr according
to the ASTM G133 standard resulted in the volume loss of 0.470 mm3 for cast samples and 0.165 mm3 for oxygen implanted
samples [147].
A generally favorable approach to improve wear properties is surface modification by application of coatings, which can
advantageously be done with tantalum. For example, a Ti-based implant component exhibited the wear rate of
1.39 ± 0.17  103 and 1.89 ± 0.4  104 mm3/Nm before and after laser application of a Ta coating, respectively [148].
The high porosity of Ta coatings supports flow of body fluids. On the other hand, it can enhance transport of wear debris
to the implant surface [31]. This phenomenon can be reduced using modern fabrication technologies. Surface modification
is also an advantageous method for improving the low wear resistance of zirconium, comparable to Ti [149]. The mass loss of
a Zr during block-on-plane testing (125 m, 1 MPa) was almost 30 mg, however, after nitrogen ion implantation it decreased
to 15 mg [149]. Wear rates of Zr bio-ceramics were already investigated as well and reported e.g. by Pandey et al. [150].
Surface treatments to improve individual properties of bio-applicable alloys are more in detail described in Section 5.
L. Kunčická et al. / Progress in Materials Science 88 (2017) 232–280 249

4.4. Fatigue

Together with wear, fatigue is among the most harmful phenomena, which can result in ultimate failures of orthopedic
implants [151]. The fatigue strength (maximal cyclic stress before the occurrence of any damage or failure) depends princi-
pally on microstructure, surface quality of the component, and loading conditions [15,151]. Fatigue failure, which usually
initiates in the most highly stressed location, is especially dangerous since it can occur before the stress level reaches the
UTS and even the YS, without any warning. Fretting fatigue, to which hip joints are especially prone, occurs when the cyclic
stress combines with the influence of friction, while fatigue supported by the effects of corrosion is referred to as corrosion
fatigue [15].
The fatigue strength of as-cast stainless steels in aqueous biological solutions is usually between 200 and 300 MPa [15].
These values are acceptable for most applications within the human body, however, they can be enhanced via subsequent
deformation processing and/or by nitrogen content increase (for Ni-free steels, preferably containing Mo). For example, the
fatigue strength of forged 316L steel was reported to reach 400 MPa [151]. Nevertheless, Co-based alloys generally have
higher fatigue strengths than stainless steels.
Owing to the fatigue strength exceeding 500 MPa, CoCr alloys are a suitable choice for applications for cyclically loaded
prostheses and any highly loaded component. However, the corrosion fatigue strength can be as low as 100 MPa for cast and
200 MPa for mechanically processed components. Modern PM technologies such as hot isostatic pressing (HIP) can
increase the fatigue strength to 700 MPa [151]. Fatigue properties can also be significantly improved by surface treatments
(see Section 5.2).
Titanium components generally feature low fatigue resistance and typically exhibit fretting fatigue failure. Fatigue prop-
erties of CP Ti and a and near–a alloys are in general not sufficient for bone-anchored joint implants, however, the fatigue
strength of a phase alloys increases with increasing oxygen content [15]. The fatigue limit of CP Ti at 107 fatigue cycles
(endurance limit) increased from 88 MPa for 0.085 wt.% of O2 to 215 MPa for 0.27 wt.% of O2. Fatigue strength and ductility
of components deteriorate also with the development of a surface a phase layer. For example, as-cast CP Ti exhibited a max-
imum fatigue stress of 325 MPa at 107 cycles, while the maximum value at 107 cycles for as-cast a” Ti7.5Mo alloy was
300 MPa. Fatigue properties were also thoroughly investigated for the most commonly used TI6Al4V alloy [152]. The fati-
gue endurance limit for components with electro polished (EP) surfaces depends on the grain size. The limits for equiaxed
and bi-modal grains were comparable (550 MPa), while they decreased for lamellar structures (350 MPa). EDM (electro
discharge machining) surface processing decreased the limit due to the deteriorated surface quality (300 MPa at 105
cycles). Papakyriacou et al. [105] reported the endurance limit of 560 MPa for Ti6Al7Nb at 2  108 cycles. The value was
by 40% lower for samples with notched surfaces compared to the polished ones, however, the notched surfaces correspond
to real conditions for stems of hip joint implants.
The endurance limit of b alloys varies between 265 and 500 MPa [15], which is lower than for stainless steels, but com-
parable to Co-based alloys. For example, the mean endurance limit for as-cast Ti13Nb13Zr was 275 MPa, although it could
have been due to the influence of interdendritic pores [153]. As reported for TNTZ, the resistance to fatigue can be increased
via Y2O3 addition [61]. The non-modified TNTZ samples exhibited a maximum cyclic stress of almost 350 MPa at 107 fatigue
cycles, while the value increased with Y2O3 additions up to 1 wt.%. The highest maximum stress of 550 MPa was observed for
TNTZ with 0.1% of Y2O3. Then, the value exhibited a decreasing trend with increasing Y2O3 content. Nevertheless, the max-
imum stress for Y2O3 additions between 0.05 and 0.5 was between 400 and 550 MPa, which was higher than for the original
TNTZ [61].
Similar to Ti, fatigue properties of Ta structures are low, however, this can be attributed to the high porosity. A Ta porous
structure prepared using SLM (80% porosity) exhibited a fatigue limit of 7.35 MPa at 106 cycles. Cold processing has only a
slight effect on fatigue properties of Ta. The reported mean endurance limit for as-cast CP Ta at 2  108 cycles and 20 kHz
was 335 MPa, while for a cold-processed CP Ta (55% reduction) it was 365 MPa. However, this value was further increased
to 410 MPa by a surface treatment (blasting and shot peening) [104].
Unlike Ta, Zr, especially in the form of bio-ceramics, is highly resistant to fatigue crack development [154]. The fatigue
limit of pure Zr subjected to reverse bending at 12 Hz was 280 MPa for 106 cycles [155], while low cycle fatigue tests
(2  104 cycles) of Zr1Nb alloy showed the fatigue strength of 300 MPa [156].

4.5. Corrosion

Corrosion is a particularly important phenomenon, since it can cause release of metal ions into the body. In vivo corrosion
properties depend on the material, mechanical and geometric properties of the implant, as well as on the surrounding envi-
ronment [20]. The degradation behavior of bio-applicable materials after being implanted is generally different than the
behavior in air [15]. The overall corrosion behavior of an implant should be evaluated with regard to other material prop-
erties, due to the possibility of development of phenomena such as stress corrosion cracking, corrosion fatigue and fretting
corrosion. The combination of corrosion with mechanical loads is especially dangerous, since it can cause e.g. a higher crack
growth rate in the vicinity of a crack initiated by pitting corrosion [157].
An important factor influencing the longevity of an implant is the pH factor, which is different in different tissues within
the human body and can temporarily change because of injury and/or inflammation and as a result of corrosion. This creates
challenging in vivo conditions for all the implanted materials [15]. Despite the fact that most metals have the ability to gen-
250 L. Kunčická et al. / Progress in Materials Science 88 (2017) 232–280

erate a passivating layer, the resistance to corrosion can generally be improved by coatings and enhancement of surfaces of
the implanted components with other elements (N2, O2) [158]. However, certain alloys with high corrosion resistances
(stainless steel) can cause allergic or inflammatory reactions, especially in long-term implantations.
Galvanic couples between dissimilar metals are the source of the most common form of corrosion [159]. Galvanic corro-
sion can be avoided in most orthopedic devices by using alloys that have passive protective oxides or that are close to each
other in the galvanic series. However, neither surface coatings, nor passivation oxide layers prevent galvanic corrosion
entirely, since pores and possible cracks in surface layers can lead to direct contact of the corrosive media with the substrate.
Moreover, hard oxide surface layers can be damaged by fretting corrosion, which leads to subsequent galvanic corrosion. For
some medical applications, non-reactive metals such as gold, silver, or platinum are used to mitigate corrosion, but these
metals have inadequate strength for orthopedics. Since galvanic couples depend on linked oxidation and reduction reactions,
one can expect that corrosion can influence biochemical reactions, including those associated with cell signaling. Galvanic
coupling can in principle influence endocrine or paracrine signaling, but this possibility has not yet been explored within
the context of orthopedic metallic implants.
Stainless steels containing at least 11 wt.% of Cr have generally good corrosion resistance, which even improves with
increasing Cr and Ni contents due to the formation of a protective oxide layer [15]. However, Cr-rich surface layers are prone
to localized degradation due to pitting and crevice corrosion [26]. Pitting corrosion is supported by the formation of Cr-rich
precipitates depleting the Cr content in their vicinity, while depletion of O2 content supports crevice corrosion, which often
occurs in locations with close contacts (e.g. between bone plates and bone screw heads). Pitting and crevice corrosion can
both be prevented by thorough fabrication via methods such as vacuum melting, which can be combined with passivation
using nitric acid performed before the final sterilization and implantation. Nevertheless, stress corrosion cracking, which is
usually initiated by the combination of tensile stress with Cl-rich corrosive environment, cannot be minimized by such pro-
gressive processing technologies. Stress corrosion failures endanger implant components containing residual stresses in
particular.
The overall corrosion behavior of austenitic stainless steels is not sufficient for long-term applications. However, Ni-free
steels with high nitrogen contents have increased resistance to crevice, intergranular and pitting corrosion [132]. Enhance-
ment of pitting corrosion resistance usually results in improvement of fatigue properties (fatigue cracks usually initiate in
the areas of pitting corrosion).
Owing to their high Cr content and the formation of a Cr2O3 surface oxide layer, the corrosion resistance of CoCr-based
alloys is far greater than for stainless steels [15]. However, the surface layer is prone to pitting corrosion in Cl-rich surround-
ings [160]. Corrosion properties of CoCr alloys can be enhanced by thermo-mechanical processing and surface treatments.
Noli et al. [158] recently documented the favorable effect of the combination of surface nitridation with subsequent oxida-
tion on improvement of CoCrMo corrosion resistance. Additions of alloying elements such as Ni and Mo improve the resis-
tance to corrosion, while additions of W deteriorate the corrosion resistance and so does the influence of friction in general
[161].
The excellent corrosion resistance of Ti is provided primarily by the protective TiO2 surface layer, however, it also depends
on the particular microstructure and chemical and phase compositions. Corrosion resistance increases substantially with
additions of elements such as Ni, Mo and Pd, and also with additions of Y2O3 [15], while it notably decreases due to the influ-
ence of fatigue. As shown by Dearnley et al. [145] for CP Ti and TI6Al4V, surface modification treatments, such as thermal
oxidation, can also significantly improve the corrosion-wear behavior. The influence of microstructure was documented
in a comparative study involving the a + b Ti6Al4V ELI samples, b-stabilized plus water quenched samples, and b-
stabilized plus furnace cooled samples. The study evaluated the corrosion behavior in Ringer’s solution. All the samples gen-
erated the passivation layer rapidly, however, the best corrosion resistance was observed for the martensitic a + b Ti6Al4V
ELI samples, the results for which were comparable to another investigated sample of b type Ti13Nb13Zr [139].
The resistance to corrosion of b phase alloys is generally slightly better than for a phase ones. This is usually caused by the
influence of noble alloying elements, which supports the generation of passivating surface oxides. For example, the Ti2448
(Ti24Nb4Zr8Sn) samples exhibited a better corrosion resistance in various physiological solutions than the CP Ti and Ti6Al4V
ones. This can be attributed primarily to the high contents of Nb2O5, TiO2, ZrO2 and SnO2 in the Ti2448 passivation layer [75].
The TiTa binary system is especially promising from the corrosion resistance viewpoint. Zhou et al. [142] investigated the
corrosion behavior of Ti plus 10, 30 and 70 wt.% Ta in Ringer’s solution. They documented a rapid generation of the resistant
surface layer consisting of TiO2 and Ta2O5 for all the samples. The corrosion resistance and stability of the surface layer
increased with increasing Ta content. The corrosion behavior of the Ti10Ta sample was comparable to the behavior of CP
Ti and Ti6Al4V.
The corrosion resistance of pure Ta is excellent, including its resistance in various aqueous solutions, chemicals and acids.
Ta is in orthopedy mainly used for protective coatings, which are usually highly resistant to corrosion. However, the corro-
sive environments have a negative effect on the mean endurance limit, which can decrease by 15% [104]. The investigation of
the corrosion behavior of Ta-coated modular necks of CoCr hip joint prostheses under dry conditions and two types of wet
conditions showed the Ta-coated components to exhibit no traces of corrosion or chemical attack [162].
Similar to Ta, the corrosion resistance of Zr is excellent. Owing to their high affinity for oxygen, Zr-based alloys develop
surface oxide layers, which increase their resistance to corrosion even more. However, the surface layer stability can still be
improved, e.g. by ions surface implantation. For example, yttrium implantation induces the generation of Y2O3 surface bar-
rier [163].
L. Kunčická et al. / Progress in Materials Science 88 (2017) 232–280 251

4.6. Compatibility with non-metallic materials

Artificial joint replacements contain many components, some of which are non-metallic, ceramic or polymeric.
Ceramic materials, mostly alumina (Al2O3) and zirconia (ZrO2), are in orthopedy mainly used for femoral heads and acetab-
ular liners, however, the use of Al2O3 has already almost been replaced with the latter [164]. Metal oxides are typically hard and
brittle, exhibit good mechanical properties, high stiffness, low wear, and low friction. This makes them a suitable choice for
components experiencing hard contacts [20,26]. Advantageous is also their high resistance to oxidation and neutral biological
response to ceramic debris particles, which enhances their overall biocompatibility [20]. On the other hand, the elastic modulus
of metal oxides is still higher than for bone. The compressive and tensile strengths of oxides are substantially different and their
pffiffiffiffiffi
fracture toughness is very low, generally less than 1 MPa m, which makes them prone to brittle fracture [151,164].
Both the mechanical and biological properties are controlled by two main factors: grain size and density, which can be
influenced through fabrication processes [20]. Small grain size and high density contribute to reduced stress, especially
the thermal stress, which is critical for ceramics, since they are not ductile and cannot compensate the external stress with
plastic deformation. The more or less inert biological behavior of ceramics is generally similar to the behavior of another
group of bio-applicable materials – polymers [20].
At present, polymers are most commonly used for articulating bearing surfaces and bone–implant inter-positional
cementing materials. Although the development of bio-applicable polymeric materials with suitable properties is rather
challenging (see Section 2.1), many have already been introduced. The polymers particularly used for human joint implants
include the following: UHMWPE, polyolefin, polydiene, silicone and HexsynTM ethylene-propylene rubbers (finger joints),
polycaprolactone (PCL) and polyactides (PLA) (knee joint menisci), acrylic resin, polyglycolides (PGA), polyhydroxybutyrates
(PHB) and, last but not least, polymethylmetacrylate (PMMA), which is typically used to fix components of large joints into
bones and thus contributes to a more uniform load distribution [20,26,165]. Since they usually experience articulating con-
tacts, polymers must exhibit a low friction coefficient and a high wear and creep fatigue resistance. The YS should also be
high enough to minimize the possibility of plastic deformation [20]. Unfortunately, the degradation rate of polymers is high
and they often cause hematomas and inflammations [166]. Although such issues have been reduced by the introduction of
hydrolytically and enzymatically biodegradable polymers, such materials are so far not suitable for joint implants.
The probably most well-known polymer in orthopedy, typically used for load bearing articulating surfaces such as tibial
spacers of knee joints and acetabular cups of hip joints, is UHMWPE [20]. Such demanding conditions usually result in a
notable wear and production of debris, however, technologies that improve wear resistance of polymers, such as increasing
of molecular weight and gamma irradiation cross-linking, have recently been introduced [167]. Nevertheless, the overall
wear rate depends on the properties of both the contacting materials. One of the contact pairs featuring a very high wear
rate (on the order of 0.1 mm/year) is UHMWPE with CoCrMo. This couple is used in large lower extremity joints, where
is usually subjected to hard-on-hard contacts; the amount of debris can reach 1  106 particles per step. The high wear rate
of joint inserts from polymers significantly decreases the longevity of the whole implant. Therefore, various contact pair
combinations have been studied to lower the wear rate as much as possible. For example, the study comparing the damage
of the polyethylene (PE) insert in the PE-CoCrMo and PE-OxiniumTM contact pairs in a knee joint implant showed a signifi-
cantly lower damage of the tibial insert for the PE-OxiniumTM contact pair [168].
Another material used in contact with metallic joint replacements is hydroxyapatite (HA), also known as calcium hydrox-
ide phosphate. HA and other similar apatites are constituents of natural bone and teeth. Its high affinity for proteins and high
compressive strength makes HA ideal for enhancing bone healing and serving as an interface to implanted metal devices.
Therefore, it often supplements metal components to support osseointegration and bone ingrowth. HA surface coatings
can be applied on metal-based joint implants to improve their biological behavior [26,169]. HA is often used in composites
owing to its excellent biological properties.
Composites combine two or more materials to obtain desirable mixtures of the properties of constituents [26]. Virtually
all composites used in medical applications are polymer-based or ceramic-based, but also include some hybrids, for example
combinations of ceramics–polymers, iron oxide–HA and Ni-free stainless steel–HA [127,170]. HA added to otherwise bioin-
ert metallic implants can impart some degree of bioactivity [171,172].
Bioactive materials can modify their surfaces after implantation. They are highly biocompatible and non–toxic to cells
and surrounding tissues, but their mechanical properties are low. Bioactive ceramics and glasses are able to connect with
the surrounding osteoblasts via direct chemical-physical bonds [20]. Bioactive calcium phosphate materials, such as hydrox-
yapatite (HA), tricalcium phosphate (TCP) and HA-based calcium phosphate ceramics (CaP), form biologically active hydrox-
ycarbonate apatite layers (HCA), which integrate with the surrounding tissues [172]. Bioactive glasses typically contain
silicon, sodium, calcium and phosphorus oxides. They undergo controlled surface degradation when in contact with physi-
ological fluids and cover themselves with a layer rich in SiO, Ca and P [26]. The originally amorphous surface layer crystal-
lizes as a mixture of hydroxycarbonate and collagen, which subsequently provides the component with bioactivity and
supports bonding of a metallic implant with the bone [20].

4.7. Biological properties

The biocompatibility of materials used in implants can be measured via assays that quantify the extents of desirable and
undesirable reactions with the human body. Biocompatible materials should be non-toxic, cause no inflammatory and aller-
252 L. Kunčická et al. / Progress in Materials Science 88 (2017) 232–280

gic reactions and should not impair functions at any level, ranging from individual cells to entire organs. The elements from
which implants are produced should in general possess a very good biocompatibility, since the biocompatibility of an ele-
ment can change depending on the exact location of implantation.
Fe and Mn are highly prone to corrosion and cause cytotoxic and genotoxic reactions [26]. Elevated Fe content in human
blood can lead to cell damage; excess Fe is stored in the heart and liver. Ni was proven to have extremely poor biocompat-
ibility, including high cytotoxicity, genotoxicity and carcinogenity, poor hemolytic behavior, potential mutagenicity and sus-
ceptibility to corrosion [173]. It can also cause severe contact allergies to skin and to the tissues surrounding implants
[15,26]. The recently developed N-rich Ni-free stainless steels possess enhanced biocompatibility, including improved cyto-
compatibility and hemocompatibility. Cr exhibits an excellent corrosion resistance and low cytotoxicity, however, Cr ions
can cause allergies and be carcinogenic and genotoxic [15]. Hexavalent Cr6+ ions are strongly oxidative and damage blood
cells, especially in the liver and kidneys. Similarly, Co has carcinogenic and cytotoxic effects and also causes allergic reactions
and hypersensitivity to the surrounding tissues. Mo is not considered as cytotoxic and its overall harmful influence on the
human body is not as severe as for the above mentioned elements (Ni, Cr, Co), but it deleteriously affects the cell proliferation
rate and mitochondrial activity [173]. W is even less toxic to living organisms, however, its corrosion resistance is weak and
loosened tungsten micro-particles are considered to be genotoxic.
Ti is generally regarded as one of the most biocompatible metals [17]. Pure Ti exhibits an excellent biocompatibility and
low cytotoxicity, both of which are attributed to the formation of a TiO2 surface layer. However, the most common alloying
elements in Ti-based alloys, V and Al, deteriorate the superior biological behavior of CP Ti. V was shown to have carcinogenic
effects, while Al has generally poor biocompatibility. It is genotoxic and can cause neurodegenerative processes and necrosis
of tissues [173]. The usage of Al in biomaterials had been controversial, since it used to be considered to cause the Alzhei-
mer’s disease, however, Al ions are not the cause for this illness, as was already proven [9]. Nevertheless, an increased
amount of Al in the body can contribute to demineralization of bones, since it competes with absorption of calcium. Al is
also harmful for kidneys and can contribute to development of women’s breast cancer [15]. Zr and Nb are both extremely
biocompatible, feature low cytotoxicity, no carcinogenity and mutagenicity, good corrosion resistance and their osteocom-
patibility is even better than for Ti [173]. Ta is considered to be one of the metals with the best biological properties. Its bio-
logical behavior is similar to the behavior of Zr and Nb [8,76,173]. Nevertheless, its oxides are suspected to be toxic to
alveoral cells [15]. Sn is considered to be non-toxic in virtually all the applications within the human organism, however,
genotoxic effects in mammalian ovary cells were reported for the SnCl2 chlorine compound [173].
Suffice to say, each of the materials used for joint replacements in orthopaedy has certain advantages and disadvantages
regarding the mechanical and corrosion properties, in vivo behavior, and compatibility with living tissue. However, most of
the properties can be enhanced by surface modification via various treatments and by applications of highly biocompatible
and resistant surface coatings.

5. Current treatment of joint alloy surfaces

The surfaces of virtually all alloys used for total hip replacement and total knee replacement in clinical practice are trea-
ted to enhance their functionality. Deliberately modifying surfaces effectively decouples the requirements placed on the bulk
properties of the joint replacement from the characteristics required only for its surface. From a certain viewpoint, all surface
treated alloys for joint applications can be viewed as bi-material composites or as functionally graded composites since they
feature advantageous properties of both the components, as illustrated in Section 4.6. Examples of surface treatments used in
currently marketed hip and knee joint replacements are listed in Table 4. The list is not intended to be comprehensive.

Table 4
Examples of surface treatments used in current hip and knee joint replacements.

Company Device Surface Treatment


Corin (UK) www.coringroup.com Trinity hip, Minihip, Cormet resurfacing, Metafix Vacuum plasma spray, Biphasic CaP electrolytic
knee, Rotaglide knee, Zenith ankle deposition
DePuy (USA) Gription Porous Ti coating, PorocoatÒ porous coating
www.depuy.com
Exactech (USA) Alteon Neck Preserving Stem Plasma spray
www.exac.com
Implants International (UK) www. Rim-fix CP Ti plasma spray
implantsinternational.com
Omni (USA) OMNIgrip-Ti, Omniknee CP Ti plasma spray, sintering
www.omniils.com
Ò
Smith & Nephew (USA) www.smith- Oxinium Surface oxidation of Zr
nephew.com
Stryker (USA) Accolade II Arc deposition, Porous Ti plasma spray, PureFixÒ
www.stryker.com LFIT HA coating, Ion implantation
Zimmer (USA) VerSysÒ EpochÒ FullCoat Hip Ti fiber metal coating, Ti porous plasma spray,
www.zimmer.com Thermal deposition
L. Kunčická et al. / Progress in Materials Science 88 (2017) 232–280 253

Instead it provides an illustration of information that implant manufacturers supply to orthopedic surgeons and medical
professionals.
Surface treatment is one of the leading means to enhance the performance of alloys for arthroplasty. The primary moti-
vations for surface treatments are to reduce wear, improve fatigue strength and enhance the biological properties. Since the
treatments to address these three motivations differ significantly, they are examined separately in the subsequent sections.

5.1. Treatments to reduce wear

Perhaps the most demanding requirement placed upon joint replacement materials is the need for their sliding interfaces
to sustain high loads for millions of cycles without degradation. This requirement has been addressed by optimizing the
combination of three interrelated engineering properties: hardness, wear resistance, and friction coefficient (Section 4.3).
Numerical models of wear have been developed which consider the complex interactions of surface properties, geometric
properties, and physiological conditions within joints [2]. Wear modeling of entire three dimensional joint systems provide
insight into how contact pressures, temperature changes, fluid film shear stresses, friction coefficients, and asperity contact
contribute to the volumetric and linear wear rate. Thus these models guide the development of materials and enable pre-
diction of how surface treatments that alter hardness, roughness, or friction may improve the overall wear resistance.
Although metal surfaces can be hardened, their ability to resist wear is intrinsically inferior to materials like ceramics or
intermetallics. Therefore, surface treatments to add ceramics or to introduce ceramic-like properties are most prominent for
reducing wear and wear debris. The results of numerical, experimental, and clinical studies also unambiguously point to the
need to alter hard metal or ceramic surfaces to reduce wear in adjacent polymers. For example, TiN coatings were applied to
Ti6Al4V and CoCrMo alloys via multiple deposition methods including Physical Vapor Deposition (PVD), Plasma Immersion
Ion Implantation (PIII), Arc Evaporative Physical Vapor Deposition (AEPVD), Powder Immersion Reactive Assisted Coating
(PIRAC), Arc Ion Plating (AIP), Arc Vapor Ion Deposition (AVID), Nitrogen Ion Implantation (NII), and Nitrogen Diffusion Har-
dening (NDH) [174]. The results of trials with these coatings varied. However, the reduction in the coefficient of friction and
reduction in volumetric wear of UHMWPE was as much as 42% in one case. In another case, the presence of pin holes or par-
tial delamination of the TiN layers increased the volumetric wear rate in UHMWPE by a factor of 4. Clinical studies of 76
patients with cemented CoCrMo hip joints with, or without, TiN coatings found loosening in 44% of implants with treated
surfaces, as compared to only 21.6% with untreated surfaces [174]. The results of six other clinical studies showed varied
results, but the evidence of TiN coating providing a significant advantage was insufficient.
Decohesion of the TiN coating and the presence of pin holes were generally identified as factors contributing to the fail-
ure. The thinness of the coating may have been a factor contributing to both pin holes and delamination. Also, the frictional
characteristics of the coating depend on the presence of lubricant between the UHMWPE and the hard surface. For example,
the presence of proteins on the surface has been shown to ameliorate the influence of surface roughness of TiN coatings
[175]. Zimmer has tried nitrogen diffusion into Ti6Al4V. As with TiN, the coating layer was too thin to last long in service.
Yildiz et al. [176,177] reported a very low wear rate for Ti6Al4V surface nitrided at 750 °C for 1 h (0.14  106 mm3/Nm, pin-
on-disc tests, 141 m). Yet other coatings have been proposed, such as ZrN for coating CoCr. Studies of the effect of implan-
tation of C and N ions on wear of the CoCrMo alloys generally show significant reduction in wear. Wang et al. [56] reported
the wear rate of surface plasma nitrided CoCrMo samples to be less than 4  108 mg/Nm at a load of 8.5 MPa. Wear rates of
CoCrMo with implanted C and N atoms were reported to reduce by a factor of 3 in ball-on-disc wear testing with an alumina
ball under the load of 5 N [178]. Ion implantation can also increase surface hardness in femoral heads, as introduced by Stry-
ker in their Low Friction Ion Treated (LFIT) CoCr heads. The depth of hardening achieved was on the order of 0.2 lm [134].

Fig. 5. Femoral stem for a total hip replacement showing the surface region roughened by titanium plasma spraying to enhance mechanical fixation during
bone ingrowth (courtesy of Exactech).
254 L. Kunčická et al. / Progress in Materials Science 88 (2017) 232–280

Coatings can also favorably decrease the wear rate of steels. The results of unidirectional sliding tests showed material losses
of 800  104, 200  104 and 2  104 lm2/cycle for uncoated 316L steel, 316L coated with 23%N nitrogen S-phase and
316L coated with CrN, respectively [133].
Wear resistance can be generally improved by reducing friction. N2Biomedical has introduced proprietary ion beam (Ion-
GlideTM) and surface coatings (SPI-ceramicTM, SPI-polymerTM) to deposit smooth metal or ceramic films to reduce surface fric-
tion and improve lubricity [179]. They deposit thin films of precious metals (Au, Ag, Pt) to serve as solid lubricants, or hard
coatings of Al2O3 or SiO2 to reduce sliding friction. Cryogenic treatments have also been shown to enhance wear resistance
by reducing friction; 196 °C treatment of Ti6Al4V increased hardness to 317.5 HV and decreased the friction coefficient to
0.20 [180]. These changes were associated with four microstructural changes – refinement of grain size, reduction of the
amount of beta phase, increase in dislocation density and increase in twin density. Mass loss measured in pin-on-disc exper-
iments was reduced by 8–19% depending upon the magnitude of applied load. The greatest improvement was obtained at
the lightest load of 50 N.
Possibly the best combination of improvement in wear resistance and reduction in friction may be achieved by surface
treatment to oxidize zirconium to form a hard ceramic surface. The OxiniumTM treatment developed by Zimmer was applied
for example to modify a Zr-2.5Nb alloy (ASTM F2384) by oxidation at 500 °C to create a hard ceramic layer [181]. Fig. 5
shows a femoral head fabricated from Zr-2.5Nb. The outer surface is oxidized via a proprietary OxiniumTM process to create
a 5–6 lm thick layer of zirconium oxide. The oxide layer is continuous with the zirconium alloy substrate, and therefore pos-
sesses several advantages over coatings. As the ZrO2 layer forms by oxidation, it expands slightly to create a compressive
surface stress, which is particularly beneficial for increasing the fatigue life. The self-grown oxide also has the advantage
of not having a distinct interface between the metal and oxide. Thus, there is no chance for debonding or failure of the
metal-ceramic interface. This is an outstanding accomplishment since there are no other examples of a metal-ceramic inter-
face surviving in a load bearing joint applications. The hardness of ZrO2 surfaces is higher than of CoCr surfaces, 12 GPa com-
pared to 5 GPa, and clinical studies have documented reduced wear rates when compared to CoCr components
[15,34,128,134,182]. The superior performance of OxiniumTM has been attributed to its lower friction and smoother surface
[128]. Data from the 2015 National Joint Registry [181] showed that only 2.67% of patients who received a Smith-Nephew
Genesis II total knee replacement needed the first revision within 10 years after the original joint replacement. This rate is
among the lowest for any implant system. In comparison, 5.39% of the patients who received the version of the Genesis II
system made with OxiniumTM needed a revision treatment within 10 years. However, the difference in performance is likely
due to the fact that the average age of the population receiving the OxiniumTM implant was much younger, 58, as compared to
an average age of 71 for the conventional Genesis II implant. The group receiving the OxiniumTM implant was likely far more
physically active, which places significantly higher demands on the implant performance. Nevertheless, additional clinical
results from comparable sample populations will be needed to document the anticipated long term benefits of OxiniumTM
implants.

5.2. Treatments to enhance fatigue strength

The classic approaches to improve fatigue strength, both of which have been implemented through surface modification,
are to reduce the stresses that initiate fatigue cracks and to reduce the presence of crack initiation sites. Peening of surfaces
by either laser or mechanical means imparts compressive stresses into surface regions, the presence of which reduces or
even eliminates the driving force for fatigue crack initiation and extension. For example, Abrasive Water Jet (AWJ) peening
applied to treat surfaces of Ti6Al4V and AISI 304 stainless steel produced compression residual stresses as high as 460 MPa in
the AISI 304 stainless steel surfaces. On the other hand, the fatigue endurance limit of Ti6Al4V increased by 25% to 845 MPa
[183]. Other mechanical surface treatments such as ultrasonic shot peening, contact rolling, and laser shock peening were
applied to alter near-surface residual stress profiles. Laser shock peening created biaxial compressive residual stresses
greater than 500 MPa to depths of 0.5 mm in Ti6Al4V [184]. This treatment used 8 ns laser pulses with an intensity of 10
GW/cm2 to generate shock pulses greater than 7 GPa over 3 mm diameter surface regions. The induced microstructural
changes were measurable to the depth of 1.6 mm and the associated increase in fatigue endurance limit was 17.2%, up to
639 MPa. The same authors also examined surface contact rolling using a 6 mm diameter hardened steel sphere rastered
across the Ti6Al4V surface under the pressure of 30 MPa. This treatment induced the compressive stresses (exceeding
1000 MPa within the first 50 lm) to the overall depth of 0.6 mm. The fatigue endurance limit increased by 13.3% to
617 MPa, which was slightly less than achieved by the above-mentioned laser shock treatment. However, the surface contact
rolling provided an additional advantage of reducing the surface roughness. It decreased the arithmetic average roughness Ra
by 64% compared to the untreated state. In examples of ultrasonic shot peening, ultrasonic energy at 20 kHz from 1 mm
diameter ceramic beads [184] and a 2.6 mm tungsten carbide ball tip [185] were applied to surfaces of Ti6Al4V. The ultra-
sonic treatment with ceramic balls increased the surface hardness from 350 HV to 400 HV to a depth of 50 lm. The corre-
sponding increase in fatigue endurance limit was from 545 MPa to 600 MPa (10.1%). Comparable increases in surface
hardness, up to 426 HV, were attained by ultrasonic surface treatments for cases in which the density of ultrasonic impacts
was greater than 70,000 strikes/mm2 [185]. X-ray and SEM analyses of ultrasonically treated Ti6Al4V surfaces revealed a
high density of twins and an increased volume fraction of b phase Ti within a highly deformed layer extending 120 lm
beneath the top surface. Sub-surface dislocation density increased from 1.25  1015 m2 to 1.85  1015 m2 after ultrasonic
treatment of 96,000 strikes/mm2. Fatigue performances for these microstructures were not reported.
L. Kunčická et al. / Progress in Materials Science 88 (2017) 232–280 255

Ion implantation not only reduces wear rates, as highlighted in Section 5.1, but also provides another means to introduce
high dislocation densities and compressive residual stresses into orthopedic alloys. Rautray et al. [5] reported improvements
on the order of 20% in fatigue life of Ti6Al4V after implantation of N and C (dosage levels of 1  1017 ions/cm2). Companies
such as N2Biomedical [179] have treated several thousand joint components from Ti6Al4V and other orthopedic alloys per
year by applying their IonguardTM ion beam process [186]. Treatments like plasma nitriding and plasma based ion implanta-
tion generate thick highly resistant surface layers of supersaturated nitrogen solutions [158]. For example, the number of
fatigue cycles to failure at the load of 4000 N for a tibial tray of a real CoCrMo knee prosthesis was 1.7  105 and
3.8  105 for the maximum fatigue stress of 250 and 400 MPa, respectively [39]
Generally, medical device manufacturers seek a 20% improvement in order to justify the adoption of experimental tech-
nologies, such as laser shock peening and abrasive water jet peening. Thus, researchers continue to probe variations or com-
binations of surface modification techniques to improve fatigue. For example, Stráský et al. [187] examined the effects of
combining electric discharge machining (EDM), etching, and shot peening to treat Ti6Al4V. Roughness induced by EDM
(see also Section 4.4) was modified by chemical etching with Kroll’s reagent and then peened with a mixture of
0.125 mm to 0.25 mm diameter balls of SiO2 and ZrO2. The EDM and chemical etching alone or in combination diminished
the fatigue endurance to less than 200 MPa, mainly due to increasing the mean surface roughness to 78 lm via EDM. How-
ever, shot peening after etching reduced the mean roughness to 15 lm, which improved the fatigue endurance limit to
450 MPa, but this fatigue strength level was still less than the 580 MPa achieved by electropolishing to reduce the surface
roughness to 1 lm. Their experiments had the additional effect of introducing microscale surface roughness to mimic struc-
ture of bone extracellular matrix to aid osteogenesis.

5.3. Treatments to enhance bone tissue integration

The interaction between joint replacement components and biological environments is largely determined by properties
of the components surfaces. Bone formation, or osseointegration, at the surfaces of replacements occurs through a series of
processes that are influenced at multiple levels by their chemical, physical, mechanical, and topological properties. These
attributes have been extensively studied [188–191].
The simplest of surface features that can be modified to enhance osseointegration is topography. Surfaces to integrate
with bone tissues are intentionally roughened to increase or accelerate osteogenesis. Surface roughness and pore sizes on
the order of 100–600 lm was determined to be the most advantageous [192–194]. Nevertheless, other attributes of surface
topography, such as pore connectivity and their influence on cellular interaction, were also studied [193]. Control of factors
such as pore throat size was demonstrated using additive manufacturing techniques, such as SLM [194]. Shot peening pro-
duces a fine scale roughness, with peak to valley heights invariably less than 50 lm, but typically on the order of 1.2–1.5 lm
[195]. Electrodeposition and sintering of powders, which are applied to roughen surfaces and to introduce other features that
promote integration of metal with the bone, are other preferred methods [31,196–202]. Plasma surface modification, which
has long been established for biomedical applications because it is particularly economical, is often applied to many implant
systems to create rough, porous surfaces that integrate well with bone [182,202–204]. It is favored especially for its ability to
produce fully dense or porous coatings. The results are readily apparent and easily understood, by surgeons and arthroscopy
patients alike [204]. An example of an Exactech tapered press-fit femoral stem, on which circumferential plasma spraying
was used to create porous titanium, is shown in Fig. 6. A clinical study of 96 patients who received this implant found that
after an average of 7.8 years, only 10% (9 implants) showed an evidence of osteolysis [205]. For comparison, the same device,
but without the surface treatment, was placed in 104 patients in the same study. Osteolysis was evident for 40% of them. This
significant difference was attributed to the benefits of the circumferential plasma spraying. In general, plasma based tech-

Fig. 6. (a) Femoral head of the Smith-Nephew Oxinium hip implant; (b) detailed view on the ZrO2 ceramic oxide film [182].
256 L. Kunčická et al. / Progress in Materials Science 88 (2017) 232–280

niques for surface modification are versatile and effective, in part because they can combine the advantages of ion beam
technologies and conventional plasma treatments.
Bioactive coatings and biomimetic treatments are among the most sophisticated approaches to enhance tissue integra-
tion for orthopedic devices. Current bioactive coatings include HA and other calcium phosphate compounds. Wang et al.
[206,207] described methods to enhance osteoconductivity with such coatings via hydrothermal deposition, plasma spray-
ing, electrochemical deposition, ion beam assisted deposition, and biomimetic deposition. Since HA is a weak and brittle
ceramic, it is usually combined with tougher bioinert ceramics such as TiO2, ZrO2, or Al2O3 to create composite layers with
suitable adhesion, hardness, mechanical strength, and bioactivity. Stryker’s Global Modular Replacement System (GMRSTM)
[208] for hip and knee replacements includes plasma sprayed coatings of porous Ti along the extra-cortical surfaces and
hydroxyapatite coatings along the proximal 30 mm of hip stems. The prospective advantages of HA coatings as compared
to porous Ti or grit blasted surfaces have been studied clinically with mixed results. Won et al. [209] found no difference
in the fixation of hip stems plasma sprayed with HA as compared to identical stems roughened by grit blasting with HA. Sim-
ilarly, Bolognesi et al. [210] found that porous HA coatings provided no clinical advantages in most cases. However, the
development of bone ingrowth was 2.6 times greater (95% confidence) for HA coated hip implants for patients with the most
severe femoral defects. Park and Lim [211] also analyzed bone ingrowth, stress-shielding, osteolysis, and complication rates
for HA coated modular hip stems and found no clinical benefits over porous Ti surfaces. Aparicio et al. [212] found that sur-
face roughening combined with alkali etching enhanced osteoblast adhesion and differentiation.
As introduced in Section 4.6, the potential of biologically active glasses as coatings have been extensively studied. They
have desirable biomechanical functions while serving as a scaffold for bone tissue formation or regeneration. [213–220].
Their application has however been limited because of their high thermal expansion coefficient (15  106 °C1) compared
to Ti alloys (9–10  106 °C1) [213]. The advantages of bioactive glasses have so far been demonstrated mainly in regener-
ation of low-load bearing bones, such as in craniofacial prostheses [219]. Techniques to apply glass coatings and achieve suit-
able microstructures and properties for load-bearing applications [221–225] are under development and will be reviewed in
Section 7 highlighting new surface treatments that are being developed toward achieving the goal of failure-free implants.
In summary, this section shows the critical importance of surface treatments for addressing the most important chal-
lenges to joint materials: reducing wear and associated debris, enhancing resistance to cyclic loads, and improving the inte-
gration of implants with endogenous bone tissue. With national registries of joint surgery outcomes showing risks of need
for revision surgery between 2% and 9%, significant room for improvement of joint alloy surfaces still remains [183]. Key
advances, along with limitations and challenges of surface treatment methods, will be summarized at the end of Section 7.

6. Advances in metals processing for joint replacements

Advances in virtually every stage of metals processing have contributed to the improvement in performance of alloys for
joint replacements. While new alloys are emerging (Section 2.2), the performance of traditional alloys continues to improve
as well by incremental improvements in production practices for metal extraction and refining, primary ingot reduction, and
secondary fabrication techniques. Advances in deformation processing technologies, including the SPD methods [44,226–
228], have enhanced the properties of metallic implant components, as have newly emerging additive manufacturing pro-
cesses. The timeliness of this review is punctuated by the recent acceleration of growth in interest in serial processing of
small metal volumes via such methods as Selective Laser Sintering (SLS) [229] and various three-dimensional printing meth-
ods (e.g. Laser Engineered Net Shaping [230], Direct Metal Printing [231], Direct Metal Laser Sintering [53]) using mixtures of
metal powders [231,232]. In these techniques, metal layers or particles are selectively heated and cooled to incrementally
build desired shapes and achieve uniform properties, or alternatively, gradients in properties. For joint replacement appli-
cations, the need for improved alloy performance must be balanced with the parallel requirement to understand the quality
and reliability of new alloys when placed into physiological environments. However, the prospect of customizing implant
alloy characteristics to patient-specific needs in nearly real-time could enable unprecedented success in implantology.

6.1. Advances in primary metal processing

Advances in extractive metallurgy and primary metal processing receive less attention since they are incremental in nat-
ure. Nevertheless, they contribute to the ever-improving success rates in the use of metals in arthroscopic devices. In general,
advances in cleanliness in alloy melting practice are critical to the overall alloy quality. Manufacturing practice enhance-
ments, such as ladle stirring and managing sources of oxygen and nitrogen during melting, have helped control inclusions,
both metallic and non-metallic [233,234]. The preparation of master alloys for making medical metals in bulk and powder
forms has benefited from more precise processing in recent years. Master alloy characteristics are particularly important in
casting and deformation processing of bulk alloys where they play very specific roles, such as in grain refinement [235,236].
They provide a cost-effective means to obtain specific chemical compositions, improve alloy homogeneity, and facilitate the
introduction of high melting temperature elements into lower temperature molten metal baths. The control of chemical
composition achieved using master alloys is a direct means to ensure reliable, homogeneous alloy properties. For example,
dispersoids (e.g. oxides, sulfides, nitrides, and carbides) to promote grain refinement during solidification or hot working can
L. Kunčická et al. / Progress in Materials Science 88 (2017) 232–280 257

Fig. 7. (a) SEM and (b) EDS maps of Fe in Grade 4 Ti showing the presence of b flecks and the associated regions of high iron concentration. The overall Fe
content of the alloy was 0.14 wt.% [own source].

be added through master alloys. Fine dispersoids serve as sites for heterogeneous nucleation. Narrow particle size distribu-
tions assure the highest rates and most uniform nucleation.
Electric Arc Furnace (EAF) melting combined with Argon Oxygen Decarburization (AOD) have improved to provide
greater homogeneity in high tonnage (>25 ton) stainless steel production, resulting in enhancement in mechanical proper-
ties [237]. Over 75% of the world’s stainless steel is made using the AOD process, which allows carbon levels to be controlled
with ±0.01 wt.% precision and to be reduced to levels less than 0.05 wt.%. Similarly, sulfur can be reduced to less than
0.001 wt.%. This level of control is particularly helpful for the low-carbon variants of austenitic stainless steels, such as
304L, 316L, and N-strengthened steels (corresponding to ASTM chemical composition requirements in standards F138,
F1314, F1586, F2229, F2581), for which the maximum carbon levels may need to be less than 0.03 wt.%, and, simultaneously,
the degree of variability for carbon in such steels can be required to be as low as 0.005 wt.% (see Section 3.1).
With the ever increasing ability to better control the melting practice, computational modeling has become one of the
most important tools for advancing the management of the complex thermodynamics and kinetics within molten steel,
which has enabled the production of cleaner stainless steels [233–235,238]. Heat-to-heat consistency achieved through
computerized process control greatly reduces the risks associated with variation in alloy quality. For example, the density
and size of non-metallic inclusions in steels (Al2O3, SiO2, TiO2), which affect their fatigue properties and deformation process-
ing characteristics, can both be reduced by optimizing the lime-alumina-based slag used in ladle refining. Modeling has been
applied to study effects of interactions between inclusions in 18Cr9Ni stainless steel melt in order to subsequently adjust the
CaO-CaF2 slag composition to reduce the content of alumina in inclusions during ladle treatments; the predictions were ver-
ified by laboratory experiments [234].
One example of titanium ingot metallurgy advancement is the improvement of vacuum arc remelting (VAR) practices,
during which Ti is melted and remelted to remove contaminants and reduce alloy segregation [239]. One defect of impor-
tance in a phase Ti alloys is b phase iron flecks which arise from segregation during VAR of ingots. The presence of b–sta-
bilizing Fe at the impurity levels allowed by the consensus standards can result in the formation of micron-sized b flecks
within the a matrix. Fe is rejected at the solid/liquid interface during solidification, resulting in macro- and micro-
segregation. This phenomenon can be controlled by introducing larger temperature gradients at the solidification interface
during VAR. Fig. 7a shows a Backscatter Electron (BSE) image of mill annealed Grade 4 (UNS R50700) Ti bar stock, which can
contain up to 0.5 wt.% Fe. The white areas of 1–1.5 lm in size are b flecks and the single dark region is a void. Fig. 7b is an X-
ray Energy Dispersive Spectroscopy (EDS) map showing variation of Fe concentration on the sample surface in the same
region scanned in Fig. 7a. The b flecks correspond with the regions of increased Fe concentration. These flecks have the desir-
able effect of increasing alloy strength. Nevertheless, they are a source of variability in alloy performance and can diminish
other properties such as corrosion, wear and fatigue resistance. The microstructures in joint replacement alloys can have
conflicting roles: sometimes enhancing specific properties, while diminishing others. This is an important theme of this
review. The multi-faceted role of b phase particles in the surgical grades of CP Ti (ASTM F67, ISO 5832-2) is one of the sim-
plest examples of the conundrum that materials scientists and engineers face in optimizing microstructures specifically for
joint replacement alloys. Tighter controls on iron impurities and on temperature gradients during VAR are now possible so
that metal producers and orthopedic device manufacturers alike can take advantage of the improved characteristics of sur-
gical grade Ti-based alloys.
Grain refinement of Ti through application of master alloys has been applied to wrought, powder, and cast forms [240–
244]. For Ti, the high solubility of most nucleant particles makes it difficult to enhance nucleation. Instead, alloy additions
that restrict grain growth are more important for reducing grain size. For example, additions of B, Si, or O can restrict grain
growth in the widely used Ti6Al4V implant alloy (ASTM F136, UNS R56401). These additions are essential since the primary
alloying elements, Al and V, do not restrict grain growth at all. Other mechanical methods such as ultrasonic treatment and
258 L. Kunčická et al. / Progress in Materials Science 88 (2017) 232–280

electromagnetic stirring are being pursued to stimulate nucleation to achieve grain refinement too [245]. Such developments
offer promising improvements in homogeneity, strength, and corrosion resistance for joint replacement alloys. These
advances should also provide synergistic benefits to the common practice of modifying replacement alloy surfaces, partic-
ularly by techniques such as ion implantation that alter sub-surface microstructures.

6.2. Advanced wrought metals processing

As is the case for primary metal processing, compelling enhancements to secondary metal processing help medical alloy
manufacturers exceed the requirements for wrought metal prostheses specified in ASTM or ISO standards. These standards
have been derived from years of experience of metal producers to rigorously document the most important process and pro-
duct characteristics. For example, standards for joint components, such as ASTM 2083-12 for Knee Replacement Prosthesis,
specify the requirements for classification, materials, performance, testing, dimensions, finishing, and packaging. The stan-
dards for materials referenced by such medical device standards provide additional requirements for joint replacement
alloys. The alloy requirements, for example, as contained in ASTM F90-09 for wrought Co-20Cr-15W-10 Ni (alloy L-605),
specify the range of chemical composition, methods of analyzing product chemistry, mechanical properties, metallurgical
requirements regarding impurities, and quality program requirements [246]. However, these standards do not specify the
microstructures that must be present, the degree of deviation from a desired microstructure that can be accepted, or what
interactions may exist between alloy microstructures and prosthesis design. In general, the standards presume that measur-
able material or device properties, appropriately inspected at the time of manufacture, will determine whether a material in
a device will perform acceptably. The current paradigm, while serving as the basis for joint alloy acceptance, can be exceeded
by a shift in focus to measuring the microstructures that evolve during manufacturing and that underlie arthroscopic device
performance.
Forming technologies have evolved to allow higher dimensional precision, better microstructure control, and creation of
more complex shapes. Simple improvements such as additional automation, improved process control and real time inspec-
tion during forming processes such as rolling, extrusion, drawing, and forging are among the leading contributors to produc-
tion of uniform microstructures. In-process quality inspection systems, such as developed by Sigma Labs Inc. (www.
sigmalabsinc.com), can be applied to monitor metallurgical variables during thermomechanical processing and to correct
possible deviations from specified process conditions in real time. Real-time inspection systems designed by metallurgists
can collect and correlate processing variables including, for example, temperatures, vibration, sound attenuation, atmo-
sphere chemistry, humidity, resistivity, capacitance, hardness, and virtually any signal for the detection of which a sensor
can be configured. Such systems help manufacturers of wrought products find and preserve the process variables that ensure
the highest performance alloys. They also document reproducibility of the conditions under which alloys are produced, and
thereby, the required microstructures and properties.
The use of three-dimensional Finite Element Analysis (FEA) has also been instrumental to optimize the forming of joint
replacement components over the past 20 years [247,248]. For the fabrication of alloy components for joints, FEA can pro-
vide information about the manufacturing process steps that is crucial for regulatory agencies (e.g., US Food and Drug
Administration, European CE mark authorities, China Food and Drug Administration) to review and approve the existing
and new medical devices. Finite element grid generation technology is evolving to enable more accurate solutions and pre-
dictions of alloy processing and performance. Discrete element size, shape, and numerical integration schemes can be
selected both automatically and manually. Complex geometries, such as commonly found in replacement joints, can be dis-
cretized automatically using algorithms to generate unstructured FEA grids [249,250]. These algorithms are designed so that
geometric non-linearities can be incorporated into mesh design. For example, steep spatial gradients in microstructures, as
represented by rapidly varying internal state variables, can be factored into automatic selection of element type and size.
Internal interfaces can also be explicitly identified and modeled.
Forging is the most common methods for shaping stainless steel, CoCr, and Ti-based alloys into hip, knee, and other pros-
thetics. While forging imparts desired geometries, it has an equally important role in modifying microstructures. The forging
process itself increases homogeneity of the microstructures and imparts strength [151,251]. One of the most significant
advances in deformation processing is the increasing use of commercial FEA codes such as DEFORM, FORGE-3D and ABAQUS
for predicting flow behavior during forming, but incorporating microstructural models to predict phase transformation, crys-
tallographic texture, grain morphology, recrystallization, grain growth, twinning, and other microstructural characteristics
that are affected by plastic flow [252–258]. Using microstructure predictions, the complex stress and temperature histories
that occur during fabrication of joint components, possibly imparting final inhomogeneous or gradient properties, can be
visualized. This capability, perhaps more than other developments in deformation processing, will enable higher reliability
and reproducibility in metallic medical implants.
Another aspect of increasingly sophisticated microstructure models is the ever-increasing level of effort needed to obtain
parameters for microstructural and phenomenological constitutive relations. For the multi-phase alloys (e.g. Ti, Cr-Co) used
in orthopedics, obtaining physically based constants for each phase is particularly challenging. Consider for example the
variety of microstructures that are present in a–b Ti-based alloys. The a phase, the properties of which alone vary dramat-
ically with alloy composition, may be acicular, plate-like, or have other morphologies. Furthermore, this phase forms either
at prior b grain boundaries or inside transformed b grains, all depending upon cooling practices. The challenge of represent-
ing such complexity is being addressed by automating the derivation of model parameters. Interpolative relationships
L. Kunčická et al. / Progress in Materials Science 88 (2017) 232–280 259

between key variables and model parameters can be developed by processing of experimental data using neural networks,
which can also be applied to determine the sensitivity of computed model predictions to variation or errors in the param-
eters [259].

6.3. Severe plastic deformation processes

Wrought alloy products experience varying degrees of deformation during their processing. When the magnitude of
imposed strains is taken to extreme levels (without annealing), particularly with the superposition of high pressure, novel
microstructural features, such as fine grain size, high angle grain boundaries, high dislocation densities, increased vacancy
concentration and refinement of the size of dispersed second phase particles, are introduced [29,44,228,260–263]. Several
fundamental physical and microstructural phenomena become important during processing of metals by SPD. Most notably,
the deformation that occurs under high pressure and concurrent high shear allows different combinations of slip, twinning,
grain boundary sliding, diffusional transport, and texture evolution. For example, during SPD by High Pressure Torsion (HPT),
pressures on the order of 2–8 GPa are imposed. Such pressures are on the order of 5% of the bulk modulus. They influence the
equilibrium concentrations of vacancies, di-vacancies, and the equilibrium width of partial dislocations during deformation
[264,265]. This alters the mechanisms by which SPD-processed metals deform both during and after processing. During SPD,
the presence of high pressure can mitigate the development of localized shear instabilities or the formation of voids. It also
reduces dislocation core widths, which enables easier cross slip, and therefore alters the dislocation substructures, and to
some extent the formation of crystallographic texture. Such effects will be most notable in joint replacement alloys with
phases possessing limited slip systems, such as the HCP low temperature a phase in titanium alloys and e phase in cobalt
alloys. Pressure during deformation can affect phase stability and transformations. In the case of Ti, sufficiently high pres-
sures during SPD can cause the formation of x phase and stabilize the structure. Similarly, the transition from the high tem-
perature a phase to the low temperature e phase in Co alloys, accompanied by a 0.15% percent increase in density, is
generally sluggish, but can be accelerated in the presence of superimposed pressure during SPD. Furthermore, finer grain
sizes in cobalt alloys stabilize the high temperature a phase at ambient temperature. Thus, different phase mixtures can
be achieved by the effect of SPD on titanium and cobalt based alloys.
The use of alloys processed by SPD in biomedical applications was recently reviewed by Lowe and Valiev [266]. They
highlighted examples in which SPD processing improved mechanical and biological properties of stainless steel, titanium,
and cobalt-based alloys. Superior mechanical properties are perhaps the most distinctive advantage of SPD-processed alloys
over their conventional coarse grain alloy counterparts. Strength increases, reported to be between 20% to over 4 times, are
accompanied by increases in low cycle and high cycle fatigue strength. Other physical and engineering properties are altered
as well, including fracture toughness, ductility, formability, corrosion resistance, thermal expansion, electrical conductivity,
and machinability. However, the associated increases in wear resistance are less substantial. This is a disappointing result
relative to the need for increased wear resistance and reduced sliding friction in joint replacement applications. Long
et al. [267,268] discussed the hardness of orthopedic alloys and the levels of wear resistance observed when subjected to
reciprocating sliding wear. They noted the formation of ultrafine grained (UFG) structures during the wear process in Ti
alloys. The high contact pressures and surface shear refine the sub-surface metal microstructure, but also create particle deb-
ris due to rupture of the oxide. In addition, stress-induced twinning and martensite formation occurred around surface wear
scratches. Thus, while the ultrafine grain structure increases surface hardness, the additional processes of oxide rupture and
reformation plus stress induced phase transformation deteriorate the frictional and wear behavior. Nevertheless, there is an
opportunity to exploit the well documented effects of SPD on diffusivity to enhance hardening of near-surface metal and
alter characteristics of surface oxides. SPD produces high densities of high angle grain boundaries with structures that
can enhance diffusivity by orders of magnitude, which has been exploited to achieve superplastic deformation even at
low temperatures [269–272]. Diffusion/oxidation surface hardening can also be enhanced. For example, diffusion of oxygen
into the Ti-13Nb-13Zr alloy produced a 2–3 lm thick diffusion hardened oxide layer [273]. An ultrafine grain size, rather
than a coarse grain size, can increase the depth of the diffusion layer, allowing even greater surface hardening. The higher
diffusivity in SPD-processed UFG alloys allows faster transport of alloying elements from their bulk into their surface oxide
layers. Metallurgists and surface engineers can exploit this characteristic to develop alloys with enhanced surface properties.
For example, Chen et al. [274] measured improved high temperature oxidation resistance of UFG 9% Cr ferritic-martensitic
steel due to enhanced diffusion of Mn. More rapid outward diffusion of Mn accelerated the formation of a thin compact sur-
face scale composed of Mn-rich oxides. This protective scale did not form on coarse-grained samples of the alloy.
Despite the fact that high pressures may also be realized during non-SPD metallurgical processes such as rolling, hot iso-
static pressing, and forward extrusion, the primary advantage of SPD methods is that shear deformation dominates over
other deformation modes to produce unique microstructures and consequent favorable properties [275]. This can be illus-
trated for example by homogenization and chemical mixing induced by shear during SPD. Ashkenazy et al. [276] performed
molecular dynamics simulation of dilute Cu alloys to examine the conditions under which shear during SPD induces various
behaviors, including chemical mixing, precipitate dissolution, and increasing the solubility of alloying elements. Ashad et al.
[277] sheared Cu90Ag10 alloys by high pressure torsion (HPT) to verify model predictions of the strain level required for
two-phase mixing of precipitates with radii between 22 nm and 107 nm. Their results confirmed that the strain to dissolve
pre-existing particles in the Cu-Ag alloy system increases linearly with particle size. These results show how shear mixing
redistributes alloy elements, enhancing the microscale homogeneity of alloy constituents. This benefit is synergistic with the
260 L. Kunčická et al. / Progress in Materials Science 88 (2017) 232–280

additional effects of having higher diffusivities in SPD-induced microstructures. The combination of enhanced homogeneity
and diffusivity is significant for joint replacement alloys which are subjected to some form of surface treatment or near-
surface modification. For example, intense shearing of alloy surfaces can be expected to have synergistic effects with ion
beam surface treatments; recovery of ion induced damage may be accelerated. Accelerating homogenization of segregated
immiscible species, which can enhance resistance to corrosion, may also be possible. The general success of SPD processing
of metals continues to propel extensive research on SPD methods, development of new SPD technologies [45,278], and
application of existing methods to new alloys and for unconventional purposes [251,266]. The combination of SPD with
PM technologies also appears to be advantageous [279].

6.4. Particle based metals fabrication

Powder metallurgy technologies enable the synthesis of alloys and attainment of properties that are difficult to achieve by
other methods [280]. Many of the benefits of conventional deformation processing to create bulk forms can be realized by
the compaction of powders. The primary purpose of PM is to fabricate solids from mixtures of elements, the blending or
melting of which aggregates the different physical properties of the particle constituents. Virtually every joint replacement
alloy can be fabricated using powder-based methods [70,242–244,281–285]. Powders produced from master alloys can be
blended and then subjected to specific combinations of temperature and pressure to achieve high levels of properties
[244,286,287]. For example, 316L stainless steel fabricated from powder was documented to exhibit a tensile strength of
300 MPa and fatigue strength of 165 MPa after sintering at 1300 °C for 30 min [288]. For very high strength alloys, bulk form-
ing methods can be so difficult that the best way to produce complex shapes is via compaction of powders [285]. Hot press-
ing of CoCr powders with diameters less than 100 lm is generally conducted at 1100–1300 °C [289]. Powder compacted
CoCr alloys possess high hardness because of the addition of 0.25–1 wt.% carbon to form carbides with Cr, W, Mo, Ta, Nb,
Zr, and Ti. These carbides become semi-coherent with the Co matrix during aging at temperatures between 700 °C and
925 °C. Hot isostatically pressing CoCrMo powders has produced fully dense alloys with a UTS of 1280 MPa [15]. This level
is only modestly less than for forged CoCrMo alloys, which have tensile strength levels between 1400 MPa and 1590 MPa.
Sintered CoCrMo components were also reported to have up to three times higher wear resistances than the cast ones
[138]. Properties of the widely used Ti6Al4V alloy can be modified via PM to approach the properties of bone tissue. Never-
theless, more research focuses on introducing and developing new Ti alloys.
Processing of powders can be more viable than processing of bulk metals and alloys especially for some difficult-to-
deform or difficult-to-cast alloys. For example, tantalum powders, which are typically applied to make porous coatings on
implant surfaces, can also be used to produce entire joint replacements components [125]. Zirconium-based alloys, such
as ZrNb or ZrNi, the fabrication of which using conventional casting technologies is very difficult, can also be preferably pro-
duced from powders [290].
The second main purpose of application of powder consolidation is to fabricate porous materials, which can in biomedi-
cine be used for porous surface coatings supporting bone ingrowth and osseointegration. However, the overall porosity and
pore size should be optimized according to the type of surrounding bone [121]. Since lower porosity could suppress regen-
eration of bone tissue and higher porosity could lead to a lower strength of newly grown bone, the pore size should be kept
between 100 and 400 lm [17,172]. The advantage of porous metals fabricated from powders is their generally lower elastic
modulus compared to fully dense solids of the same composition produced by casting. However, porosity lowers strength
[291].
The production of powders itself often involves ball milling, which imparts large deformations into the powders, causing
significant grain refinement [242]. Another example of powder processing is spray forming or spray casting. The size and
shape distribution of powders can be controlled to create reproducibly porous structures on surfaces, such as in total hip
replacement systems. This control of powder characteristics imparts superior mechanical properties and corrosion
resistance.

6.5. Additive manufacturing processes

Serial manufacturing has emerged as a potentially instrumental technology for biomedical metals and alloys. It offers
extraordinary opportunities for patient care, enabling customization of the geometries and microstructures in medical
devices to meet the unique physiological needs of individuals. Additive Manufacturing (AM) provides a distinctly new char-
acteristic for medical device manufacturing industry – the ability to create superior material systems, but without the com-
mon practices of incorporating proprietary alloy formulations. Trademarked joint replacement alloy formulations such as
TMZFÒ [292], Trabecular MetalÒ [125], and OxiniumTM [34] emerged from long periods of development, medical testing,
and optimization of manufacturing and associated quality control practices. With the adoption of serial 3D metal printing,
alloy composition can be varied during fabrication of individual component. Functional gradients in microstructures and
properties can be created which are derived from in-process mixing of master alloyed powders. Time-temperature histories
for cooling of each incremental volume of alloy can be made unique. Thus both the degree of microstructure control and
complexity of fabrication increase together.
A necessary precursor for the use of AM technologies in medical devices is the ASTM standards. Such standards have
emerged for Additive Layer Manufacturing (ALM) of Ti6Al4V (ASTM F2924) and Ti6Al4V ELI (ASTM F3001), and for related
L. Kunčická et al. / Progress in Materials Science 88 (2017) 232–280 261

Fig. 8. Metal human vertebra made using a Computer Tomography (CT) model via Selective Laser Melting (SLM) [293,381].

powder-based approaches, such as Metal Injection Molding in ASTM F2989 and ASTM F2885 for CP Ti and Ti6Al4V, respec-
tively [293]. The standards for AM components specify both product requirements and process controls. This contrasts with
the traditional standards for bulk materials which instead focus on controlling the properties.
Several labels have emerged as synonyms for overlapping categories of the leading serial materials fabrication technolo-
gies: Additive Manufacturing (AM) [294], 3D Printing (3DP) [295], Rapid Prototyping (RP) [296], and Solid Freeform Fabri-
cation (SFF) [229]. Specific subclasses of these serial manufacturing techniques include Additive Layer Manufacturing
(ALM) [297], Laser Based Metal Deposition (LBMD) [298], Direct Metal Laser Sintering (DMLS) [238], Selective Laser Sintering
(SLS) [229], Selective Laser Melting (SLM) [299], Electron Beam Freeform Fabrication (EBFF) [300], and Electron Beam Melt-
ing (EBM) [301]. Proprietary names include Laser Engineered Net Shaping (LENSÒ) and Directed Light Fabrication (DLFÒ). In
general, additive approaches to metal fabrication include laser based and electron beam based manufacturing approaches,
but related older technologies, which also employ powders, are concurrently gaining momentum, for example Metal Injec-
tion Molding (MIM), Conventional Hot Pressing (CHP), Inductive Hot Pressing (IHP), Cold Isostatic Pressing (CIP), Spark
Plasma Sintering (SPS), and Direct Powder Rolling (DPR) [302]. In these techniques, the granularity of the metal manufactur-
ing is defined by the size scale of the feedstock (multi-micron sized powders) or the volume of material being serially pro-
cessed (cubic microns to cubic millimeters). Geometric properties and microstructural characteristics are concurrently
determined in AM, as opposed to conventional bulk metal processing methods in which alloy producers create the most uni-
form possible starting stock, and the geometry is determined subsequently by subtracting metal by machining, stamping, or
shaping.
Fig. 8 shows an example of a human vertebra produced by SLM. Such techniques demonstrate the ability to readily repro-
duce complex geometries. However, the level of properties, including biological ones, continues to be investigated. Investi-
gations of mechanical properties generally have shown adequate performance for Ti-based alloys [293,297,303,304],
stainless steel [305,306] and Co-based alloys [307,308]. The biological responses of serially manufactured Ti-based alloys
processed e.g. by SLS, EBM, and MIM have been evaluated [293,381]. However, the residual porosity inherent to sintering
may generally present a risk of reduced fatigue, fracture, and corrosion properties. For example, SLS processed CoCr alloys
promoted bone growth significantly more than conventionally cast CoCr parts [307]. CoCr components processed by SLS also
featured good mechanical properties, but with residual porosity, which was speculated to diminish corrosion resistance
[308].
As is commonly the case with powder processed materials, the entire suite of engineering properties depends upon clean-
liness of the starting powders and the degree of densification achieved. As could be expected, phase compositions, textures,
and microstructures of laser processed materials can be significantly different from conventionally processed alloys of nom-
inally the same compositions. Thus, to some extent alloys produced from powders can be regarded as fundamentally new
and different from their conventional counterparts. For example, Ti-Nb-Zr-Ta (TNZT) b alloys produced via the LENSÒ process
exhibit a homogeneous b phase with nanometer-scale precipitates, including x phase [309]. However, the amount of oxygen,
a significant a phase stabilizer and strengthener, is significantly less than in conventional TNZT. LENSÒ deposited TNZT also
forms a h0 0 1i solidification texture, which in turn influences the tendency for localized shear-banding and the anisotropy of
mechanical properties [310]. This anisotropy may explain the higher tensile strength and yield strengths compared to con-
ventional TNZT.
In summary, this section highlights the recent evolution of primary metal production techniques. Incremental improve-
ments in foundry melting practices, impurity control, and computational modeling have converged to make available highly
homogeneous and reliable bulk medical alloys. The combination of advances in real-time inspection technologies and
262 L. Kunčická et al. / Progress in Materials Science 88 (2017) 232–280

automated in-process controls has further honed alloy production quality. New methods of controlling microstructure via
Severe Plastic Deformation have provided bulk alloys of unprecedented strength, inspiring worldwide interest in further
developing SPD methods for improving medical metals. The primary advantage of SPD methods is the ability to impart ultra-
fine internal structure features within bulk metals. Most recently, powder based fabrication has expanded to encompass 3D
additive manufacturing processes. The early success of selective powder fusing and solidification into complex shapes is
driving rapid development to exploit the advantages of 3D printing of metals and explore the limits of geometric and com-
positional complexity that can be achieved.

7. Advances in treating metallic joint surfaces

As discussed in Section 5, the most direct drivers for advancing surface treatments in alloys for arthroplasty have been to
minimize osteolysis associated with wear debris and to alter the biological or metallurgical processes that enhance integra-
tion of alloy surfaces with bone tissues. Surface alterations to reduce wear are applied to sliding interfaces while the later are
applied to regions of bone contact, such as hip stems. In both categories, the surfaces are invariably non-planar, and may
have intricate geometries, concavities, and possess fine scale features. These characteristics constrain the options for surface
treatment technologies and have prompted the development of multiple novel approaches for joint surface modifications.
From a scientific viewpoint, surfaces possess the complexity of being interfaces between distinct materials, phases, or
environments. Even without any treatment they are intrinsically planar discontinuities which, at the microscale, may also
have sophisticated three dimensional characteristics. Consequently, applying surface engineered treatments to alter the
properties of regions of matter that are already hybrids between materials with differing electronic and physical states
remains among the most challenging areas of theoretical and experimental materials science. The high level of complexity
of the technical challenges of advancing surface science and surface modification processes increases the need for stronger
theoretical underpinnings. Today, the state of theoretical knowledge of surfaces at the atomic scale is still emerging, and is
based on a compilation of competing semi-empirical and first principles quantum mechanical models. Even for pure metals,
the predictions of theory provide mainly qualitative insights, but with ever improving quantitative predictive accuracy. For
example, using empirical n-body Finnis-Sinclair potentials, the range of equilibrium surface energies (at 0 K) for clean, unre-
constructed metal surfaces has been predicted to be between 0.62 J/m2 for a silver (1 1 1) surface to 3.036 J/m2 for a (3 1 0)
surface on tungsten [311]. However, application of a semi-empirical model derived from specific surface energies measured
at the point of melting yielded a different result: the highest surface energy of a metal (tungsten) was computed to be 3.25 J/m2
at 0 K [311]. Considering that top-surface and sub-surface phenomena occurring in metals are driven largely by reduction in
surface free energies, the need for quantitative accuracy becomes evident. The understanding of chemisorption, physisorp-
tion, adhesion, and surface segregation of low energy species into the high energy surface regions of alloys, or their complex
oxides, depends on the ability of researchers to quantify surface energies, both theoretically and experimentally. Thus, the
advances in the theory of surfaces provide an important backdrop to the advances emerging in surface treatment.
From a purely technological viewpoint, surface treatment also offers multiple challenges. Surface scientists and engineers
must master the intricacies of the alloy substrate, understanding how modifications of only the surface region can be
achieved to enhance local properties, while preserving the desired bulk alloy properties. It is their responsibility to under-
stand, with equal depth of knowledge, the materials processing and characteristics of joint replacement alloys (e.g. Ti, Fe,
CoCrMo, Ta, and Zr) while modifying just the surface regions to meet the needs for specific medical applications. Moreover,
all surface processes invariably include gradients of composition and field variables such as temperature, stress, and elec-
tronic potential.
For most biomedical applications, advancements in surface treatment are also further challenged by the fundamental
commandment to which health care providers adhere, to ‘‘do no harm” to their patients. The potential for such harm has
emerged repeatedly, for example in the recall of the DePuy ASR hip replacement in 2010, the safety alert issued for Biomet’s

-/+
hν m axby

cz
(a) (b) (c) (d)
Fig. 9. Four categories of surface alteration for joint replacement alloys (a) sub-surface modification without elemental addition, (b) sub-surface
modification with addition of one or more elements m, (c) top-surface modification without elemental addition, and (d) top-surface addition of layers by
deposition of elemental or molecular compounds axby, or growth of layers by transport of elemental constituent cz [own source].
L. Kunčická et al. / Progress in Materials Science 88 (2017) 232–280 263

Table 5
Categories and examples of surface treatment techniques.

Category Technique Examples


Sub-surface Heating Furnace treatment, flame hardening, convective, and conductive heating
modification Electromagnetic irradiation Laser treatment, ultraviolet treatment, gamma sterilization, inductive heating,
without adding magnetic stirring & hysteresis
elemental species Ionizing irradiation Gamma, beta, positron, electron beam, neutron beam, proton beam
Mechanical modification Shot peening, grit blasting, ultrasonic peening, Friction stir processing
Sub-surface addition Beam Ion beam, Cluster Ion Beam, Molecular Beam, Accelerated Neutral Atom Beam
by implanting
elemental species
Plasma Plasma Sputtering, Plasma Immersion Ion Implantation, High Intensity Plasma Ion
Nitriding,
Diffusion Carburizing, Nitriding, Nitrocarburizing, Carbonitriding, Boriding, Boronizing
Vacuum CVD, PVD, Ion Plating Plasma Assisted
Top-surface Material removal Chemical etching, spark erosion
modification Surface patterning, roughening, or Mechanical patterning, ion milling, photolithography, electron beam lithography,
chemical activation chemical etching, broad band visible light
Top-surface additions Conversion coating Anodizing, black oxide coating, chromate coating, phosphate coating, stannate
by depositing or coating, cerium oxide coating
growing layers Spray High Velocity Oxygen Fuel spray coating, Flame spraying, Plasma Spraying
Chemical Nanophase patterning, nanodot arrays

(now Zimmer’s) M2a Magnum, Stanmore, and Exceed ABT hip joint replacements in 2010 and 2012, and the Stryker volun-
tary recall of its Rejuvenate Modular and ABG II in 2012 [312]. These are all products containing metal-on-metal sliding
interfaces, carefully designed to exceed the performance of the prior generation of metal-on-polymer based joints. Yet in
2013 the UK banned all metal-on-metal hip implant designs based on the high failure rates associated with metallosis doc-
umented in the UK national patient database [181]. These occurrences underscore the need for the technical community to
further advance materials engineering, especially surface engineering, to improve how materials function in joint replace-
ment devices.
It is helpful to differentiate existing and prospective new surface treatment methodologies into four categories: (a) sub-
surface modification without adding elemental species, (b) sub-surface addition of elemental species, such as ions, clusters of
atoms, or compounds, (c) top-surface modification without adding elemental species, and (d) top-surface addition of ele-
mental species by depositing or growing new surface layers. These categories are illustrated in Fig. 9 and exemplified by
techniques listed in Table 5. Sub-surface modification includes adding energy via heat, light, atomic particle bombardment,
and electromagnetic radiation. These methods produce gradients in structural defects and stored energy in the surface
region. The second category, sub-surface implantation, includes adding elemental or molecular constituents into alloy sub-
strates, with depths ranging from 10 nm to several millimeters. This category may involve ballistic injection of negatively
charged masses m- or positively charged masses m+ into the alloy matrix. Such methods alter the chemical free energy, stres-
ses, and chemical composition of the surface region. The third category, top surface modification, includes patterning,
reshaping, and removal of material via macroscale mechanical bombardment, chemical etching, or other mechanical means.
The last category, top surface addition, involves deposition of atomic or molecular compounds axby, in uniform or non-
uniform layers. Some methods of surface treatment concurrently introduce effects from more than one of these four cate-
gories. For example, plasma treatments subject surfaces to a mixture of ions, electrons, molecules, and electromagnetic
fields. It is important to understand the synergistic influences of different mechanisms acting simultaneously upon alloy sur-
faces in order to control surface treatments to achieve desired properties.
Although alternative schemes for categorizing surface modification methods have been proposed [204,313], viewing sur-
face techniques for joint replacement alloys in these categories is helpful from multiple perspectives, including from the
viewpoints of materials scientists, materials process modelers, biomedical engineers, medical device manufacturers, and reg-
ulatory agencies. The scheme depicted in Fig. 9 helps delineate the different natures of sub-surface vs. top-surface engineer-
ing, and the differences between modifying pre-existing alloy constituents vs. adding new ones.
The ingenuity of researchers propelling the science and engineering of surfaces has generated an exhaustive array of sur-
face treatment methods, the scope of which is beyond what can be addressed in this review. Instead, selected developments
in each of the four categories that appear most promising for advancing arthroplasty are highlighted. Section 5 reviewed sev-
eral surface treatment technologies that are at the forefront of current treatment practices, and have been implemented in
clinical studies. In contrast, this section examines techniques that may have been tested in vitro but have yet to appear in
clinical studies. Another contrast is that the most advanced techniques sometimes combine two or more methods of surface
treatment. The trend towards application of multiple methods is a result of the fact that no single technique has completely
resolved the current challenges facing arthroplasty alloys yet. Furthermore, the combined effects of multiple techniques are
clearly additive, and even synergistic.
264 L. Kunčická et al. / Progress in Materials Science 88 (2017) 232–280

Fig. 10. Schematic depiction of large strain machining showing cutting depth t, rake angle a, and chip thickness tc [322].

7.1. Sub-surface modification

Subsurface modification techniques alter the intrinsic properties of the alloy in the region within a millimeter of its free
surface. Thus, this category of treatment depends largely on the metallurgical characteristics of the alloy. Very significant

Fig. 11. EBSD grain orientation maps for copper machined under four conditions: (a) 0° rake angle and cutting speed of 50 mm/s, (b) 0° rake angle and
cutting speed of 550 mm/s, (d) 20° rake angle and cutting speed of 50 mm/s, and (e) 20° rake angle and cutting speed of 550 mm/s [321].
L. Kunčická et al. / Progress in Materials Science 88 (2017) 232–280 265

changes in properties such as hardness or strength are possible for alloys that respond to heating and cooling transients (heat
treatable steels). Localized heating followed by rapid cooling have well understood effects in causing phase transformations
and introducing non-equilibrium microstructures. One example is the common martensite formation in steels with carbon
contents between 0.35 wt.% and 0.5 wt.%. However, these levels of carbon are far higher than found in most stainless steels
used in orthopedics (see Section 3). Consequently, localized heat treatments are rarely applied for Fe-based joint materials.
As discussed in Section 3, Ti-based and Co-based alloys exhibit a range of microstructures that depend sensitively upon
cooling rate, and include amorphous and metastable phases [314,315]. The cooling rates required to achieve some of these
microstructures and phase morphologies are difficult to achieve in as-fabricated joint replacement components, but can be
attained in powders heated by Selective Laser Melting (SLM) and cooled during AM (Section 6.5) [314]. Thus, the classical
notion of surface treatments to achieve local surface properties readily extends to surfaces of the small volumes of alloys
that can be selectively added e.g. via AM technologies. These opportunities, stimulated in part by the first successful surgical
placement of 3D printed hip joints, are just beginning to be explored [316,317].
Among the most promising subsurface alteration techniques for joint replacement components are mechanical treat-
ments that impart deformation-induced microstructural changes. They are simple to implement and result in high densities
of planar defects such as twins and grain boundaries. For example, the Severe Shot Peening (SSP) [318] method and Surface
Mechanical Attrition Treatment (SMAT) [319], both providing grain refinement, have been shown to create nanocrystalline
structures in surfaces [320]. SSP of AISI 316L stainless steel has been found to increase Vickers microhardness from the nom-
inal level of 170 HV to over 380 HV, enhance strength, compressive residual stresses, osteoblast adhesion and proliferation,
while it decreased the growth of gram-positive bacteria without the presence of antibiotics [318]. Another means to severely
deform surfaces to create desirable microstructures is machining. The process of material removal by orthogonal metal cut-
ting imparts large strain and high shear into the near-surface regions [321]. Generally, the microstructures within the top
50 lm of a machined surface are significantly altered [322]. Fig. 10 illustrates the region in front of an orthogonal cutting
tool that is modified by machining. Although the largest strains occur in the chip that is removed, the surface below the plane
of material separation from the bulk experiences significant strains as well.
Shekhar et al. [321] analyzed the distribution of strains, strain rates, and temperatures experienced during machining at
various rake angles and speeds. They expressed their results in terms of the Zener-Holloman parameter (Eq. (1)):
 
Q
Z ¼ e_ exp ð1Þ
RT
where e_ is the strain rate, Q is the activation energy, R is the gas constant, and T is the absolute temperature. The variation
in Z within the microstructurally affected zones during machining is on the order of 12 orders of magnitude. Changing the
cutting speed or rake angle of the cutting tool causes microstructure gradients beneath the surface to vary, as shown in
Fig. 11 for machining of copper. Maps of gradients in Z can be used together with technical data to predict microstructures
and associated variations in hardness and strength.
Other variants of mechanical sub-surface modification include Modulation Assisted Machining (MAM) and Large Strain
Extrusion Machining (LSEM). In both methods, techniques designed originally for mechanical surface removal have been
morphed into microstructure modification schemes. In MAM, low frequency (<1000 Hz) oscillation is applied to orthogonal
cutting tools to create nanostructured and UGF alloys with high strength and hardness [322–325]. The process can create
spherical, platelet, needle, and fiber shaped particles with unique deformation induced microstructures. These particles
can subsequently be consolidated onto surfaces or used to fabricate entire joint replacements. Using LSEM, difficult to
deform alloys such as Inconel 718 were nanostructured, demonstrating the suitability of the technique to other high strength
alloy systems, including titanium and stainless steel [323].
Mechanical treatments can be readily combined with other techniques such as ion implantation and electromagnetic and
ionizing radiation, which can add heat and introduce subsurface defects. The depth of these effects is limited to sub-micron
dimensions, determined by the incident energy and atomic particle fluence. Light-based or c-irradiation treatments were
applied to bio-polymers and composites with bioactive glasses [214], but have not been used with alloys for joint replace-
ment yet. However, with the emergence of additive manufacturing methods for 3-dimensional printing, the prospect of cre-
ating multi-materials (e.g. www.stratasys.com) that are composites of metals, ceramics, and polymers, is realizable. Low
energy or low fluence surface treatments may be used to modify the polymer component of metal-polymer hybrids, inducing
curing or other chemical changes in near-surface regions.

7.2. Sub-surface addition

7.2.1. Ion implantation


Ion implantation is among the most viable surface modification techniques for joint replacement alloys due to its partic-
ularly large effects on wear resistance [5,326–328]. In conventional implantation processes, ions are accelerated inside a vac-
uum chamber to energies (typically between 10 keV and 200 keV) sufficient for the ions to penetrate the surface of the
target alloy. However, penetration depths are usually less than 1 lm, since the ions lose energy through a cascade of elastic
collisions with the atomic nuclei, inelastic collisions with the electrons of surface atoms, and charge exchange processes
[5,326]. Elastic collisions, causing atom displacements from their equilibrium positions, sometimes imparting amorphization
266 L. Kunčická et al. / Progress in Materials Science 88 (2017) 232–280

Fig. 12. Hardness profile for carburized AISI 316 stainless steel after (a) gas carburizing at 420 °C; (b) gas nitriding at 445 °C [339].

or other forms of damage to the crystal structure, dominate at low initial ion energies, while inelastic collisions dominate at
higher energies and incident ion velocities. Defects caused by elastic collisions may quickly recover at elevated temperatures,
or may be intentionally removed by post-implantation annealing. Ion collisions produce distinct changes in surface
microstructures, which can be controlled to tailor surface properties. Substrates for conventional ion implantation are most
often heat treated to enable recovery of defects.
Methods of conventional ion implantation, along with more advanced techniques such as Low Energy-High Temperature
Ion Implantation (LEHT), Plasma Immersion Ion Implantation (PIII), and Plasma Source Ion Implantation (PSII) have been
overviewed by Garcia and Rodriguez [329]. PSII offers multiple advantages over conventional ion implantation techniques,
particularly for implantation of nitrogen ions in medical alloys [330]. In PSII, a plasma sheath surrounds the target surface,
eliminating the need for line-of-sight configurations. PSII avoids the need for beam rastering, target surface masking, and
enables cost-effective high throughout processing of large surface areas.
Since wear rates of the hardest and most wear resistant CoCrNo alloys in sliding interfaces are still inferior to e.g. ceramic-
metal and ceramic-ceramic interfaces, their surfaces are excellent candidates for further enhancement by ion implantation
[135]. Garcia et al. demonstrated improvements in tribological behavior of a Co28Cr8Mo alloy subjected to an incident dose
of 3  1018–6  1018 O2 atoms/cm2 via PIII [331]. The PIII process increased the surface hardness by more than 80%, reduced
the coefficient of friction by more than three times, and decreased as much as ten times the wear rate after oxygen was
implanted at the highest temperature studied (610 °C). Similar improvements in wear resistance were reported for CoCr
implanted with nitrogen [134,186,332–334]. Ion implantation reduced wear rates (measured by weight loss) to levels typ-
ically found in sliding wear of zirconia [186]. Werner et al. implanted both C and N into CoCrMo with total fluences up to
7  1017 ions/cm2 at temperatures between 200 °C and 400 °C [178]. Wear rate decreased by a factor of 3, with no evidence
of the formation of any carbide phases, CrN or Cr2N, and the depth of nitrogen penetration increased from 0.4 lm at 200 °C to
3 lm at 400 °C. Ion beam treatments were also applied to Ti alloys [5,24,203,335,336] and stainless steels [333]. For Ti, the
formation of TiN and Ti2N increased the hardness and wear resistance [24], while for stainless steel, carbon implantation was
also found to induce nanoscale precipitations [337].

7.2.2. Thermochemical diffusion methods


Diffusional based methods for hardening surfaces are best known for treating the surfaces of steels, for which the addi-
tions of C and N2 via gas or liquid media can increase surface hardness to 70 Rc [338]. These methods create case depths
between 0.010 and 0.200 mm, with the microhardness nearest the surface in the range of 800–1600 HV. Low temperature
diffusion treatments can be applied to all stainless steels, and have been successfully applied to duplex and precipitation
hardening stainless steel grades. Like carburizing, nitriding introduces interstitials into the austenite lattice, providing sur-
face hardening to depths of 10 lm, as shown for two levels of nitrogen potentials KN in Fig. 12 [339]. Both carburizing and
nitriding impart compressive stresses into the surfaces of stainless steel and can be combined to increase the depth of the
hardened region to 40 lm. This has the double beneficial effect of improving wear resistance and increasing the fatigue
endurance limit. Low temperature salt bath treatments of stainless steel impart hard layers with thicknesses up to 50 lm
featuring hardnesses up to 1200 HV, and two orders of magnitude improvement in wear resistance (as measured by volume
of wear debris) [340]. For AISI 304 and 316 austenitic steels, nitriding can produce surface hardnesses as high as 1600 HV
[341].
Carbon-based treatments of stainless steel surfaces can however deplete Cr from the surface region through the forma-
tion of Cr23C6 carbides. Therefore, they reduce the amount of Cr available to form the protective Cr2O3 oxide that is so impor-
tant for providing corrosion resistance for all stainless steels (see Section 3.1). In general, any process that heats stainless
steel into the temperature range between 600 °C and 800 °C enables carbon diffusion to form chromium carbides, and
thereby sensitizes the steel. Gas nitriding is typically performed at lower temperatures, between 500 °C and 550 °C, below
the sensitization temperature range for stainless steel. Nitrogen treatments can however deplete Cr if conducted at elevated
L. Kunčická et al. / Progress in Materials Science 88 (2017) 232–280 267

temperatures (600–800 °C) at which CrN and Cr2N readily form [342]. Fortunately, treatments with carbon at temperatures
below 450 °C can be applied to increase surface hardness to 1000–1200 HV due to the addition of interstitial carbon without
carbon reacting with Cr [339]. Similarly, treatments with nitrogen can be conducted at reduced temperatures, between
300 °C and 450 °C without significant depletion of Cr from the steel matrix. However, Manova et al. [343] found nanoscale
CrN precipitates at grain boundaries in TEM foils of 304 stainless steel treated at 350 °C. In subsequent investigations, Lutz
et al. [344] found no CrN formation in medical CoCr (Alloy L605) processed at 350 °C, but nevertheless noted surface passi-
vation was diminished. They hypothesized that the affinity of chromium for nitrogen reduced the mobility of chromium in
the alloy, thereby prohibiting the formation of chromium oxide to passivate the surface.
Adding nitrogen to CoCrMo alloys has been demonstrated by multiple methods. Nitriding at an elevated temperature
introduces multiple precipitates into the surface region, including Cr2N and p phase. Reactive diffusion coatings have been
applied via Powder Immersion Reaction Assisted Coating (PIRAC) to Zr, Ti and CoCrMo surfaces to achieve extraordinary high
wear resistance in clinical trials [345]. Surfaces of PIRAC-treated hip joints retrieved from dogs after 3.5 years showed no
visible evidence of wear. Plasma nitriding of CoCrMo (ISO5832-12) formed CrN and Cr2N, which can increase the hardness
up to 2740 HV and 2175 HV, respectively [342]. Layers of CrN and Cr2N up to 14 lm thick formed in the a-CoCr matrix, and
resulted in wear rate reduction by a factor of 4.
To titanium and its alloys, carburizing and nitriding can be applied via pulsed plasma treatment, gas treatment, and high
frequency induction heating. For induction heating, nitride layers with depths of 30 lm have been reported, with hardnesses
of 2000 HV [346]. Slightly lower depths (25 lm) and hardnesses (1780 HV) can be achieved by introducing carbide into tita-
nium via high frequency induction heating. These treatments improve wear resistance due to the formation of hard phases
including TiN, Ti2N, and the presence of interstitial nitrogen.

7.3. Top-surface modification

One of the most biologically significant characteristics of abiotic materials in biotic environments is their topography. For
the sliding surfaces between the materials in joint replacements the elimination of surface roughness is essential, whereas at
interfaces between alloys and bone tissue, an optimum rough surface topography is sought to maximize bone ingrowth or
other aspects of tissue interaction. Top-surface modifications to create rough topographies rank among the best approaches
to achieve bone integration. Similarly, surface polishing by various means is an essential aspect of surface preparation for
sliding interfaces.
Surface topographies can be created through addition or removal of material. All the methods that add matter possess the
intrinsic issue of creating an interface, including the prospect of debonding. In contrast, patterning or removal of bulk metal
from a surface avoids the risk of interfacial delamination. Surface topographies range from stochastic to highly symmetric 2D
lattices. Stochastic surfaces can be created by grit blasting, chemical milling or etching, while patterning can be achieved
through lithographic methods. Park et al. [347] showed how irregular surface indentations made by grit blasting on titanium,
which produced arithmetic average surface roughness Ra between 1.48 lm and 1.51 lm with maximum peak-to-valley
heights between 16.09 lm and 18.41 lm, increased surface energy and wettability, and thereby enhancing osteoblast cell
adhesion, viability and differentiation. Nanoprint lithography has been used to create pillars of varying heights ranging from
2 lm to 12 lm and spacings ranging from 6 lm to 50 lm to examine the effect of patterning on protein uptake, which cor-
related with increased surface area, but only when the patterned surfaces were also treated with oxygen plasma to make
them hydrophilic [348]. A mechanism for how surface topography affects protein adsorption was postulated by Scopelliti
et al. [349] in terms of how roughness influences protein-protein interactions to form multilayers. Others have hypothesized
that proteins bind inside valleys of grooved surfaces [350,351]. These mechanisms may operate to varying degrees in the
formation and ongoing stabilization of joint alloy/bone tissue interfaces, whether on coated or uncoated surfaces, or on fully
dense or macroscopically porous surfaces. For example, Stryker uses 72% dense TritaniumÒ in acetabular shells with a large
pore size of 546 lm. Coatings of HA can also be applied, for example, Stryker’s PureFix coating which adds 50 lm of ceramic
to create a macro-scale rough surface [208].
On the other hand, atomistically smooth surfaces are most desirable for sliding interfaces in joints. Mechanical, chemical,
or electrolytic polishing can create nanometrically smooth alloys surfaces, which reduce the propensity for abrasive wear in
all metal-based material couples: Metal-on-Metal, Metal-on-Polymer, and Metal-on-Ceramic [352–355]. In abrasive wear,
surface asperities cut into the opposing surfaces causing scratching and generating debris (Section 4.3). However, adhesive
wear occurs when a local chemical bond melds contacting surfaces until a segment of one of the surfaces is pulled away leav-
ing a pit or a scratch. Adhesive wear is considered to be the predominant wear mechanism in joint replacement implants
[119]. This mechanism occurs even between atomistically smooth surfaces. Avoidance of adhesive wear can be accomplished
more effectively through top-surface additions.

7.4. Top-surface addition

Coatings on the top surfaces of joint replacement alloys have the distinct characteristic of isolating the bulk substrate
alloy from interacting chemically with the surrounding material. This is in part why this approach attracts vigorous research
interest. However, it also introduces additional challenges, most notably adhesion of the top layer to the substrate and avoid-
ance of local damage to the layer that may promote localized corrosion. The elusive goal of eliminating adhesive wear
268 L. Kunčická et al. / Progress in Materials Science 88 (2017) 232–280

demands the addition of ceramic-like surfaces, which have intrinsic bonding characteristic that mitigate chemical melding.
One particularly promising approach is the addition of graphene nanoplatelets as reinforcements to alumina ceramic layers.
Nieto et al. [356] showed that adding 5–15 vol.% of graphene nanoplatelets to alumina increased the wear resistance by 39%
and fracture toughness by 21%. Diamond-like coatings have also been investigated, but with concerns about adhesion [357–
359], while composites of Ti and C-60 fullerenes have been examined as both micropatterned and continuous coatings [360].
Ongoing investigation of such composites is merited, especially to achieve desirable combinations of tribocorrosion resis-
tance and biocompatibility.
In some cases, improving corrosion resistance is a central motivation for top-surface additions, although the alloys used in
joint replacement form adherent, passivating oxides providing general corrosion resistance and ensuring minimal release of
metal ions into the surrounding tissues. Nevertheless, elevated levels of Cr, Ni, Co, Ti, V, and Al have been found in the liver,
spleen, and other tissues of total joint arthroplasty patients [135]. Co levels in the liver of normal humans have been mea-
sured to be 120 lg/g, while levels of 15,200 lg/g were measured for healthy patients with normally functioning implants.
These high post-surgery levels of metal ions are associated with their uptake from fretting, which results in more readily
dissolved wear debris (Section 4.3). Fine debris particles can activate human immune responses leading to inflammation.
Specific pattern recognition receptors in the membranes and the cytosol of immune responses cells such as macrophages
induce secretion of proinflammatory cytokines [135]. Implant debris ingested by macrophages may subsequently destabilize
lysosomes causing the release of ingested intracellular metal ion contents. Thus, surface treatments that minimize wear and
associated debris formation are critically important for all articulated joints.
A potential to enhance wear and corrosion resistance may be realized by the creation of metallic glasses on the surfaces of
implant components. Metallic glasses may allow customization of the surface elastic properties too. Abdi et al. [361] studied
TiZrNiSi glass forming alloys produced via melt-spinning, comparing the oxide film growth with conventional surfaces of
commercial purity a-titanium and a b-Ti-40Nb alloy. Although they found that the corrosion rates in simulated body fluid
and the apatite forming ability of the glassy alloy were similar to untreated conventional alloys, the glasses had the advan-
tage of possessing lower elastic moduli and higher wear resistance. Nevertheless, additional research is needed to make
bioactive glass technology useful for joint applications.
The most successful use of top-surface coating technology remains the addition of porous structures to surfaces that inte-
grate with bone tissue [362]. A wide range of methods have been used successfully including conventional powder metal-
lurgy, thermal plasma spraying, chemical vapor deposition, selective laser sintering, micro-arc oxidation, and rapid
prototyping [363–365]. Micro-arc oxidation used in combination with superhydrophobic surface chemical treatment has
been shown to simultaneously enhance corrosion resistance and in vitro hemocompatibility of Ti6Al4V surfaces [365]. Such
combinations of surface treatment methods appear to offer the greatest promise for useful advances in joint alloy surface
properties, despite their additional complexity.
In summary, this section examines surface treatments that have the potential to replace or extend the current state-of-
the-art treatments identified in Section 5. A scheme for categorizing treatment approaches is proposed that differentiates
between top-surface and sub-surface modifications and between methods that add elemental matter or not.
The distinction between top-surface and sub-surface treatments differentiates between modifications for which interface
chemistry and adhesion are of the greatest importance (top surface treatments) rather than gradient effects (sub-surface
treatments). Top surface additions, such as diamond-like coatings, introduce the benefits of extreme wear or corrosion resis-
tance, but at the expense of creating sharp interfaces prone to decohesion. In contrast, sub-surface treatments create near-
surface gradients avoiding interfacial discontinuities and debonding. Gradient engineering can impart exceptional properties
to surfaces regions. It can also be applied more broadly to engineer bulk material properties. The sub-surface effects of ion
implantation and thermochemical diffusion on hardness and wear resistance both rely on the ability of the substrate to host
high concentrations of added elements and introduced structural defects. Diffusional methods are intrinsically more cost
effective, but can only be applied to substrates that can survive elevated temperatures without unwanted side effects. Con-
sequently, the frontiers of development of diffusional schemes press toward lower temperature treatments, while ion
implantation methods evolve toward lower cost non-line-of-sight treatments.
The distinction between incorporating additional elements or only modifying the already present substrate material
highlights the importance of thermodynamic variables. Adding elements to the top-surface or sub-surface increases the
complexity of the thermodynamic system chemically and physically. For example, small atoms of carbon or nitrogen alter
the chemical free energy of surface regions, and introduce the possibility of reaction with the substrate to form additional
phases. Additional mechanical energy is also stored in the strain fields surrounding the introduced defects. Since elemental
additions create diffusion couples, implanted species may interact with intrinsic substrate species and alter the kinetics of
their transport. Overall, elemental additions impart complexity to surfaces beyond non-elemental surface medication tech-
niques. This points to the importance of ever improving theoretical and computational models to understand the effects of
elemental additions and design surface treatment schemes.

8. Outlooks and horizons

The advances in metals processing and surface treatment that were highlighted in the prior sections have all contributed
to the ever-improving standard of care available for joint replacements. In this section, some horizons in materials science
L. Kunčická et al. / Progress in Materials Science 88 (2017) 232–280 269

and the extremes of properties that we believe may be achievable, with sufficient resources and research, are briefly men-
tioned together with promising 3D imaging and non-destructive examination technologies. The attention is focused on those
areas that today appear to be of the greatest impact on joint arthroplasty.

8.1. Horizons in materials

New materials will emerge as additive manufacturing technologies evolve. Selective deposition and fusing of particles
introduces the prospect of melding minute volumes of polymers, ceramics, and metals to create composites in situ. This
prompts the notion of integrated material processing, by which the techniques to concurrently process multiple classes
of materials give rise to multi-materials. While the early embodiment of multi-material 3D printing involves only blending
of multiple polymers (e.g. www.stratasys.com), the concept can be extended to particles with more diverse properties. Com-
binations of elemental metallic powders, polymer stocks, and ceramic particles can be integrated to produce gradients
within medical devices that would otherwise be difficult or impossible to achieve. For example, interleaving bone-like
ceramics with porous metallic structures could provide unprecedented integrations of bone tissues and metals. These pos-
sibilities are just beginning to be explored.
Extremes in the ability of alloys used in joint replacement to sustain high loads, resist wear, and environmental degra-
dation will depend on the ability to create hybrids. Even in single phase metals, the highest levels of strength are achieved
through the action of multiple mechanisms, such as ultrafine grain sizes combined with the presence of nanoscale twinning
[366,367]. The traditional emphasis on creating homogeneous simple bulk alloys has been consistent with the high value
placed upon reliability and reproducibility in the performance of medical devices. However, this emphasis has reached
the limit of its utility for joint replacement applications. New material classes may evolve from unusual combinations of
materials that are intrinsically inhomogeneous and complex. For example, covetic metals can contain over 20 at.% carbon
organized into islands of graphene-like layers or networks of carbon nanoparticles [368,369]. Other hybrid materials have
been shown to significantly enhance strength and wear resistance, for example composites containing silicon carbide nano-
whiskers and nanoparticles embedded in epoxy [370]. The greatest success in pushing the limits of material performance
will most likely be found in materials systems such as these that are intrinsically complex. Fortunately, the availability of
improved characterization methods, fine spatially discrete serial fabrication methods, and more powerful computational
models of material properties together make such complex materials feasible. With careful development and systematic lab-
oratory and clinical evaluation, some of these emerging materials technologies may be optimized for joint replacement alloy
applications.

8.2. Horizons in technologies

Advances in arthroscopic alloys depend in part on ancillary technologies that support advances in materials science.
Three-dimensional materials science has become one of the most significant enablers of progress with new materials. The
growth of 3D imaging technology, new characterization technologies, and rapid prototyping techniques each merit mention
here since they support the development of new classes of materials. For example, X-ray, neutron, and proton based tomog-
raphy along with 3D atom probe technology are allowing unprecedented insight into the internal structure of metals, with
ever improving spatial resolution. X-ray ultra-microscopy methods such as Transmission X-ray Microscopy are able to image
features down to 30 nm [371]. While the highest resolution micro computed tomography techniques are restricted to ex vivo
applications, new methods for High Resolution Peripheral Quantitative Computed Tomography (HR-pQCT) have been used to
image bone microstructures in the tibia in vivo. Proton and neutron techniques enable deeper penetration into alloys, but
with a lower resolving power. Detectors for proton radiography and tomography have only recently been designed and
tested in clinical applications [372]. Such techniques can be expected to evolve to be used in conjunction with other probes,
such as acoustics to non-invasively quantify the performance of bioengineered joint surfaces and surrounding tissues [373].
One emerging acoustic characterization technique for examining joint materials is Resonant Ultrasound Spectroscopy
(RUS). RUS provides a precise means to sample the thermodynamic properties of materials by measurement of elastic con-
stants [374,375]. The tensor components the elastic modulus can be probed by causing small elastic displacements of atoms
about their equilibrium positions in a crystalline solid. The moduli are spatial derivatives of the free energy, and as such,
provide information about multiple physical properties, including specific heat, Debye temperature, phase transitions,
and anharmonicity [376]. A wide range of complex shapes can be analyzed by lightly contacting the sample surface at just
two points between transducers, or in one RUS instrumentation configuration employing an electromagnetic acoustic trans-
ducer (EMAT), with no contact of the sample at all. RUS has already been used in medical imaging to identify early micro-
damage in cortical bone [377], to determine the relationship between bone damping and density in the tibia [378], to
establish the effects of hydration on anisotropy of human dentin [379], and to measure the elastic response of open-cell por-
ous medical alloys used in the femoral stem of a total hip joint replacement and in the tibial plate of a total knee joint
replacement [362]. Very small variations in the values of the elastic modulus and in the degree of viscous damping of elastic
waves can be detected and correlated with changes in tissue quality and properties of collagen-mineral compounds near
bone-implant interfaces. Additional ultrasound technologies have been proposed for characterization, for example, by pho-
toacoustic imaging in vivo and sonochemical synthesis of nanostructures in vitro [373].
270 L. Kunčická et al. / Progress in Materials Science 88 (2017) 232–280

9. Conclusion

Joint replacement alloys have been used over the past 50 years with significant success. Millions of people have received
arthroplastic surgeries extending their mobility and quality-of-life by years. Nevertheless, significant opportunities remain
to improve joint replacements by improving the materials from which they are made. While metals, ceramics, and polymers
each serve critical functions in implant systems, their respective roles and properties are shifting with the advancements in
each category and the introduction of hybrid material systems.
We have summarized the status of alloy technology currently in use, and have shown the following:

 Titanium, stainless steel, and CoCr remain the most extensively incorporated alloys in joint replacement. Enhancements
to each of these alloy systems continue to be explored, as well as new alloys based on alternative metals such as Zr and Ta.
Within these alloy families, new methods to produce fully dense and porous microstructures have emerged, ranging from
increasingly sophisticated melt practices with master alloys, to deformation-induced homogenization by shear mixing, to
final construction of three dimensional shapes by powder based laser deposition and spark plasma sintering.
 The mechanical, corrosion, biological, and tribological properties of the principle classes of alloys have persistently
improved. Yet greater improvements have resulted from modifications of medical device surfaces by a wide variety of
localized treatments and elemental additions. Combinations of multiple surface modification approaches are proving
to be most effective. For example, sub-surface additions by ion beam implantation in combination with top-surface mod-
ifications by nanophase patterning are revealing synergistic effects that have established record high tribocorrosion
performance.
 Hybrids of metals, ceramics, and intermetallic compounds appear most promising. For example, the internal oxidation of
zirconium to create gradated hard surface layers of zirconium oxide or electric arc treatments to create graphene-like lay-
ered regions in metals exemplify some of the novel processes and microstructures that are being investigated.
 The ever-improving abilities to image, quantify, and model microstructures using new instrumentation and computation-
intensive methods are augmenting the tool kits that materials scientists and surface scientists can apply to joint alloy
development. Three-dimensional measurement and tomographic imaging is leading the way to freeform fabrication of
complex shapes with tailored gradient microstructures.

Finally, no single advancement has dominated the ongoing improvements in joint replacement alloys. Rather, it has been
the evolution of theory, computational modeling, and a combination of experimental technologies that have been system-
atically evaluated in vitro and in vivo which have, and will continue to, advance the entire field. The ultimate benefactor of
these spiraling synergistic improvements is the human patient who can experience longer lived, trouble free implants.

Acknowledgement

This paper was created in the Project No. LO1203 ‘‘Regional Materials Science and Technology Centre - Feasibility Pro-
gram” funded by Ministry of Education, Youth and Sports of the Czech Republic and through the support of the George S.
Ansell Department of Metallurgical and Materials Engineering Project Development Fund at the Colorado School of Mines.

References

[1] Shahadat M, Teng TT, Rafatullah M, Arshad M. Titanium-based nanocomposite materials: a review of recent advances and perspectives. Colloids Surf B
Biointerfaces 2014;126:121–37. doi: http://dx.doi.org/10.1016/j.colsurfb.2014.11.049.
[2] Mattei L, Di Puccio F, Piccigallo B, Ciulli E. Lubrication and wear modelling of artificial hip joints: a review. Tribol Int 2011;44:532–49. doi: http://dx.
doi.org/10.1016/j.triboint.2010.06.010.
[3] Geetha M, Singh AK, Asokamani R, Gogia AK. Ti based biomaterials, the ultimate choice for orthopaedic implants – a review. Prog Mater Sci
2009;54:397–425. doi: http://dx.doi.org/10.1016/j.pmatsci.2008.06.004.
[4] Guillemot F. Recent advances in the design of titanium alloys for orthopedic applications. Expert Rev Med Devices 2005;2:741–8. doi: http://dx.doi.
org/10.1586/17434440.2.6.741.
[5] Rautray TR, Narayanan R, Kim K-H. Ion implantation of titanium based biomaterials. Prog Mater Sci 2011;56:1137–77. doi: http://dx.doi.org/10.1016/
j.pmatsci.2011.03.002.
[6] Antunes RA, de Oliveira MCL. Corrosion fatigue of biomedical metallic alloys: mechanisms and mitigation. Acta Biomater 2012;8:937–62. doi: http://
dx.doi.org/10.1016/j.actbio.2011.09.012.
[7] Pezzotti G, Yamamoto K. Advances in artificial joint materials. J Mech Behav Biomed Mater 2014;31:1–2. doi: http://dx.doi.org/10.1016/j.
jmbbm.2013.12.012.
[8] Abdel-Hady Gepreel M, Niinomi M. Biocompatibility of Ti-alloys for long-term implantation. J Mech Behav Biomed Mater 2013;20:407–15. doi:
http://dx.doi.org/10.1016/j.jmbbm.2012.11.014.
[9] Niinomi M, Nakai M, Hieda J. Development of new metallic alloys for biomedical applications. Acta Biomater 2012;8:3888–903. doi: http://dx.doi.org/
10.1016/j.actbio.2012.06.037.
[10] Voisin M, Ball M, O’Connell C, Sherlock R. Osteoblasts response to microstructured and nanostructured polyimide film, processed by the use of silica
bead microlenses. Nanomed-Nanotechnol Biol Med 2010;6:35–43. doi: http://dx.doi.org/10.1016/j.nano.2009.05.007.
[11] Goodman SB, Barrena EG, Takagi M, Konttinen YT. Biocompatibility of total joint replacements: a review. J Biomed Mater Res A 2009;90:603–18.
[12] Williams DF. Implantable prostheses. Phys Med Biol 1980;25:611–36.
[13] Hänzi AC, Gerber I, Schinhammer M, Löffler JF, Uggowitzer PJ. On the in vitro and in vivo degradation performance and biological response of new
biodegradable Mg-Y-Zn alloys. Acta Biomater 2010;6:1824–33. doi: http://dx.doi.org/10.1016/j.actbio.2009.10.008.
[14] Nedoma J, Stehlík J, Hlaváček I, Daněk J, Dostálová T, Přečková P. Mathematical and computational methods in biomechanics of human skeletal
systems. John Wiley & Sons, Inc.; 2011.
L. Kunčická et al. / Progress in Materials Science 88 (2017) 232–280 271

[15] Chen Q, Thouas GA. Metallic implant biomaterials. Mater Sci Eng R 2015;87:1–57. doi: http://dx.doi.org/10.1016/j.mser.2014.10.001.
[16] Eynon-Lewis NJ, Ferry D, Pearse MF. Themistocles Gluck: an unrecognised genius. BMJ 1992;305:1534–6. doi: http://dx.doi.org/10.1136/
bmj.305.6868.1534.
[17] Oldani C, Dominguez A. Titanium as a biomaterial for implants. In: Fokter SK, editor. Recent Adv Arthroplast. InTech; 2012. p. 149–62. doi: http://dx.
doi.org/10.5772/1445.
[18] Song E, Seon J, Moon J, Ji-Hyoun Y. The evolution of modern total knee prostheses. In: Arthroplast – Updat. InTech; 2013. p. 184–95. doi:http://dx.doi.
org/10.5772/54343.
[19] Bonnin M, Amendola A, Bellemans J, MacDonald S, Ménétrey J. The Knee Joint. Paris: Springer Paris; 2012. doi: http://dx.doi.org/10.1007/978-2-287-
99353-4.
[20] Hallab J, Jacobs JJ. Biomaterials science: an introduction to materials in medicine. In: Buddy D. Ratner, Allan S. Hoffman, Frederick J. Schoen JEL,
editors. 3rd ed. Oxford: Elsevier Academic Press; 2013.
[21] Kocich R, Szurman I, Kursa M, Fiala J. Investigation of influence of preparation and heat treatment on deformation behaviour of the alloy NiTi after
ECAE. Mater Sci Eng A 2009;512:100–4. doi: http://dx.doi.org/10.1016/j.msea.2009.01.054.
[22] Thomasová M, Seiner H, Sedlák P, Frost M, Ševčík M, Szurman I, et al. Evolution of macroscopic elastic moduli of martensitic polycrystalline NiTi and
NiTiCu shape memory alloys with pseudoplastic straining. Acta Mater 2017;123:146–56. doi: http://dx.doi.org/10.1016/j.actamat.2016.10.024.
[23] Szurman I, Kocich R, Kursa M. Shape memory alloys - fabrication and processing. Saarbrücken: LAP Lambert Academic Publishing GmbH & Co. KG;
2012.
[24] Lanning BR, Wei R. High intensity plasma ion nitriding of orthopedic materials. Surf Coat Technol 2004;186:314–9. doi: http://dx.doi.org/10.1016/
j.surfcoat.2004.02.047.
[25] Wan P, Ren Y, Zhang B, Yang K. Effect of nitrogen on blood compatibility of nickel-free high nitrogen stainless steel for biomaterial. Mater Sci Eng C
2010;30:1183–9. doi: http://dx.doi.org/10.1016/j.msec.2010.06.015.
[26] Talha M, Behera CK, Sinha OP. A review on nickel-free nitrogen containing austenitic stainless steels for biomedical applications. Mater Sci Eng C
2013;33:3563–75. doi: http://dx.doi.org/10.1016/j.msec.2013.06.002.
[27] Xiong Y, Qian C, Sun J. Fabrication of porous titanium implants by three-dimensional printing and sintering at different temperatures. Dent Mater J
2012;31:815–20. doi: http://dx.doi.org/10.4012/dmj.2012-065.
[28] Valiev RZ, Semenova IP, Latysh VV, Rack H, Lowe TC, Petruzelka J, et al. Nanostructured titanium for biomedical applications. Adv Eng Mater 2008;10:
B15–7. doi: http://dx.doi.org/10.1002/adem.200800026.
[29] Lukac P, Kocich R, Greger M, Padalka O, Szaraz Z. Microstructure of AZ31 and AZ61 Mg alloys prepared by rolling and ECAP. Kov Mater Mater
2007;45:115–20.
[30] Vandamme NS, Que L, Topoleski LDT. Carbide surface coating of Co-Cr-Mo implant alloys by a microwave plasma-assisted reaction. J Mater Sci
1999;34:3525–31. doi: http://dx.doi.org/10.1023/A:1004670207579.
[31] Levine BR, Sporer S, Poggie Ra, Della Valle CJ, Jacobs JJ. Experimental and clinical performance of porous tantalum in orthopedic surgery. Biomaterials
2006;27:4671–81. doi: http://dx.doi.org/10.1016/j.biomaterials.2006.04.041.
[32] de Souza KA, Robin A. Preparation and characterization of Ti–Ta alloys for application in corrosive media. Mater Lett 2003;57:3010–6. doi: http://dx.
doi.org/10.1016/S0167-577X(02)01422-2.
[33] Watari F, Yokoyama A, Omori M, Hirai T, Kondo H, Uo M, et al. Biocompatibility of materials and development to functionally graded implant for bio-
medical application. Compos Sci Technol 2004;64:893–908. doi: http://dx.doi.org/10.1016/j.compscitech.2003.09.005.
[34] Hernigou P, Mathieu G, Poignard a, Manicom O, Filippini P, Demoura A. Oxinium, a new alternative femoral bearing surface option for hip
replacement. Eur J Orthop Surg {&} Traumatol 2007;17:243–6. doi: http://dx.doi.org/10.1007/s00590-006-0180-2.
[35] Castellani C, Lindtner RA, Hausbrandt P, Tschegg E, Stanzl-Tschegg SE, Zanoni G, et al. Bone-implant interface strength and osseointegration:
biodegradable magnesium alloy versus standard titanium control. Acta Biomater 2011;7:432–40. doi: http://dx.doi.org/10.1016/j.actbio.2010.08.020.
[36] Kocich R, Kunčická L, Dohnalík D, Macháčková A, Šofer M. Cold rotary swaging of a tungsten heavy alloy: numerical and experimental investigations.
Int J Refract Met Hard Mater 2016;61:264–72. doi: http://dx.doi.org/10.1016/j.ijrmhm.2016.10.005.
[37] Kocich R, Kunčická L, Mihola M, Skotnicová K. Numerical and experimental analysis of twist channel angular pressing (TCAP) as a SPD process. Mater
Sci Eng A 2013;563:86–94. doi: http://dx.doi.org/10.1016/j.msea.2012.11.047.
[38] Piccinini M, Cugnoni J, Botsis J, Ammann P, Wiskott A. Numerical prediction of peri-implant bone adaptation: comparison of mechanical stimuli and
sensitivity to modeling parameters. Med Eng Phys 2016;38:1348–59. doi: http://dx.doi.org/10.1016/j.medengphy.2016.08.008.
[39] Villa T, Migliavacca F, Gastaldi D, Colombo M, Pietrabissa R. Contact stresses and fatigue life in a knee prosthesis: comparison between in vitro
measurements and computational simulations. J Biomech 2004;37:45–53. doi: http://dx.doi.org/10.1016/S0021-9290(03)00255-0.
[40] Savoldelli C, Bouchard P-O, Loudad R, Baque P, Tillier Y. Stress distribution in the temporo-mandibular joint discs during jaw closing: a high-
resolution three-dimensional finite-element model analysis. Surg Radiol Anat 2012;34:405–13. doi: http://dx.doi.org/10.1007/s00276-011-0917-4.
[41] Amirouche F, Choi KW, Goldstein WM, Gonzalez MH, Broviak S. Finite element analysis of resurfacing depth and obliquity on patella stress and
stability in TKA. J Arthroplasty 2013;28:978–84. doi: http://dx.doi.org/10.1016/j.arth.2013.02.002.
[42] Halloran JP, Petrella AJ, Rullkoetter PJ. Explicit finite element modeling of total knee replacement mechanics. J Biomech 2005;38:323–31. doi: http://
dx.doi.org/10.1016/j.jbiomech.2004.02.046.
[43] Baldwin Ma, Clary CW, Fitzpatrick CK, Deacy JS, Maletsky LP, Rullkoetter PJ. Dynamic finite element knee simulation for evaluation of knee
replacement mechanics. J Biomech 2012;45:474–83. doi: http://dx.doi.org/10.1016/j.jbiomech.2011.11.052.
[44] Kocich R, Macháčková A, Kunčická L. Twist channel multi-angular pressing (TCMAP) as a new SPD process: numerical and experimental study. Mater
Sci Eng A 2014;612:445–55. doi: http://dx.doi.org/10.1016/j.msea.2014.06.079.
[45] Kocich R, Kunčická L, Macháčková A. Twist Channel Multi-Angular Pressing (TCMAP) as a method for increasing the efficiency of SPD. IOP Conf Ser
Mater Sci Eng 2014;63:12006. doi: http://dx.doi.org/10.1088/1757-899X/63/1/012006.
[46] Zach L, Kunčická L, Růžička P, Kocich R. Design, analysis and verification of a knee joint oncological prosthesis finite element model. Comput Biol Med
2014;54:53–60. doi: http://dx.doi.org/10.1016/j.compbiomed.2014.08.021.
[47] Ingrassia T, Nalbone L, Nigrelli V, Tumino D, Ricotta V. Finite element analysis of two total knee joint prostheses. Int J Interact Des Manuf
2012;7:91–101. doi: http://dx.doi.org/10.1007/s12008-012-0167-7.
[48] Bei Y, Fregly BJ. Multibody dynamic simulation of knee contact mechanics. Med Eng Phys 2004;26:777–89. doi: http://dx.doi.org/10.1016/j.
medengphy.2004.07.004.
[49] Kraaij G, Zadpoor Aa, Tuijthof GJM, Dankelman J, Nelissen RGHH, Valstar ER. Mechanical properties of human bone-implant interface tissue in
aseptically loose hip implants. J Mech Behav Biomed Mater 2014;38:59–68. doi: http://dx.doi.org/10.1016/j.jmbbm.2014.06.010.
[50] Ganesan V, Laha K, Mathew MDD. Influence of nitrogen content on the evolution of creep damage in 316 LN stainless steel. Proc Eng 2014;86:58–65.
doi: http://dx.doi.org/10.1016/j.proeng.2014.11.011.
[51] Ma Z, Zhao X, Ge C, Ding T, Li J, Xiao X. A new resource-saving, high chromium and manganese super duplex stainless steels 29Cr-12Mn-2Ni-1Mo-xN.
J Iron Steel Res Int 2011;18:41–6. doi: http://dx.doi.org/10.1016/S1006-706X(12)60032-6.
[52] Casabán Julián L, Igual Muñoz A. Influence of microstructure of HC CoCrMo biomedical alloys on the corrosion and wear behaviour in simulated body
fluids. Tribol Int 2011;44:318–29. doi: http://dx.doi.org/10.1016/j.triboint.2010.10.033.
[53] Barucca G, Santecchia E, Majni G, Girardin E, Bassoli E, Denti L, et al. Structural characterization of biomedical Co–Cr–Mo components produced by
direct metal laser sintering. Mater Sci Eng C 2015;48:263–9. doi: http://dx.doi.org/10.1016/j.msec.2014.12.009.
[54] Liao Y, Pourzal R, Stemmer P, Wimmer MA, Jacobs JJ, Fischer A, et al. New insights into hard phases of CoCrMo metal-on-metal hip replacements. J
Mech Behav Biomed Mater 2012;12:39–49. doi: http://dx.doi.org/10.1016/j.jmbbm.2012.03.013.
272 L. Kunčická et al. / Progress in Materials Science 88 (2017) 232–280

[55] Wang Q, Zhang L, Shen H. Microstructure analysis of plasma nitrided cast/forged CoCrMo alloys. Surf Coat Technol 2010;205:2654–60. doi: http://dx.
doi.org/10.1016/j.surfcoat.2010.10.031.
[56] Wang Q, Zhang L, Dong J. Effects of plasma nitriding on microstructure and tribological properties of CoCrMo alloy implant materials. J Bionic Eng
2010;7:337–44. doi: http://dx.doi.org/10.1016/S1672-6529(10)60265-X.
[57] Tipper JL, Firkins PJ, Ingham E, Fisher J, Stone MH, Farrar R. Quantitative analysis of the wear and wear debris from low and high carbon content cobalt
chrome alloys used in metal on metal total hip replacements. J Mater Sci Mater Med 1999;10:353–62.
[58] Scholes SC, Unsworth A. Pin-on-plate studies on the effect of rotation on the wear of metal-on-metal samples. J Mater Sci Mater Med
2001;12:299–303. doi: http://dx.doi.org/10.1023/A:1011238902803.
[59] Dobbs HS, Robertson JLM. Heat treatment of cast Co-Cr-Mo for orthopaedic implant use. J Mater Sci 1983;18:391–401. doi: http://dx.doi.org/10.1007/
BF00560627.
[60] Regener B, Krempaszky C, Werner E, Stockinger M. Modelling the micromorphology of heat treated Ti6Al4V forgings by means of spatial tessellations
feasible for FEM analyses of microscale residual stresses. Comput Mater Sci 2012;52:77–81. doi: http://dx.doi.org/10.1016/j.commatsci.2011.03.035.
[61] Song X, Niinomi M, Nakai M, Tsutsumi H, Wang L. Improvement in fatigue strength while keeping low Young’s modulus of a beta-type titanium alloy
through yttrium oxide dispersion. Mater Sci Eng C 2012;32:542–9. doi: http://dx.doi.org/10.1016/j.msec.2011.12.007.
[62] Nakai M, Niinomi M, Hieda J, Yilmazer H, Todaka Y. Heterogeneous grain refinement of biomedical Ti–29Nb–13Ta–4.6Zr alloy through high-pressure
torsion. Sci Iran 2013;20:1–4. doi: http://dx.doi.org/10.1016/j.scient.2013.01.004.
[63] Gao Z, Li Q, He F, Huang Y, Wan Y. Mechanical modulation and bioactive surface modification of porous Ti–10Mo alloy for bone implants. Mater Des
2012;42:13–20. doi: http://dx.doi.org/10.1016/j.matdes.2012.05.041.
[64] Li YH, Chen RB, Qi GX, Wang ZT, Deng ZY. Powder sintering of porous Ti-15Mo alloy from TiH2 and Mo powders. J Alloys Compd 2009;485:215–8. doi:
http://dx.doi.org/10.1016/j.jallcom.2009.06.003.
[65] Zhou YL, Niinomi M, Akahori T. Effects of Ta content on Young’s modulus and tensile properties of binary Ti–Ta alloys for biomedical applications.
Mater Sci Eng A 2004;371:283–90. doi: http://dx.doi.org/10.1016/j.msea.2003.12.011.
[66] Zhou YL, Niinomi M, Akahori T. Decomposition of martensite a00 during aging treatments and resulting mechanical properties of TiTa alloys. Mater
Sci Eng A 2004;384:92–101. doi: http://dx.doi.org/10.1016/j.msea.2004.05.084.
[67] Xu L, Xiao S, Tian J, Chen Y. Microstructure, mechanical properties and dry wear resistance of b -type Ti  15Mo  x Nb alloys for biomedical
applications. Trans Nonferrous Met Soc China 2013;23:692–8. doi: http://dx.doi.org/10.1016/S1003-6326(13)62518-2.
[68] Li C, Zhan Y, Jiang W. Beta-Type Ti-Mo-Si ternary alloys designed for biomedical applications. Mater Des 2012;34:479–82. doi: http://dx.doi.org/
10.1016/j.matdes.2011.08.012.
[69] Miura K, Yamada N, Hanada S, Jung TK, Itoi E. The bone tissue compatibility of a new Ti-Nb-Sn alloy with a low Young’s modulus. Acta Biomater
2011;7:2320–6. doi: http://dx.doi.org/10.1016/j.actbio.2011.02.008.
[70] Málek J, Hnilica FF, Veselý J, Smola B. Heat treatment and mechanical properties of powder metallurgy processed Ti-35.5Nb-5.7Ta beta-titanium alloy.
Mater Charact 2013;84:225–31. doi: http://dx.doi.org/10.1016/j.matchar.2013.08.006.
[71] Málek J, Hnilica FF, Veselý J, Smola B, Bartáková S, Vaněk J, et al. The influence of chemical composition and thermo-mechanical treatment on Ti-Nb-
Ta-Zr alloys. Mater Des 2012;35:731–40. doi: http://dx.doi.org/10.1016/j.matdes.2011.10.030.
[72] Yilmazer H, Niinomi M, Nakai M, Cho K, Hieda J, Todaka Y, et al. Mechanical properties of a medical b-type titanium alloy with specific microstructural
evolution through high-pressure torsion. Mater Sci Eng C Mater Biol Appl 2013;33:2499–507. doi: http://dx.doi.org/10.1016/j.msec.2013.01.056.
[73] Raducanu D, Vasilescu E, Cojocaru VD, Cinca I, Drob P, Vasilescu C, et al. Mechanical and corrosion resistance of a new nanostructured Ti – Zr – Ta – Nb
alloy. J Mech Behav Biomed Mater 2011;4:1421–30. doi: http://dx.doi.org/10.1016/j.jmbbm.2011.05.012.
[74] Zhang DC, Yang S, Wei M, Mao YF, Tan CG, Lin JG. Effect of Sn addition on the microstructure and superelasticity in Ti-Nb-Mo-Sn Alloys. J Mech Behav
Biomed Mater 2012;13:156–65. doi: http://dx.doi.org/10.1016/j.jmbbm.2012.04.017.
[75] Bai Y, Hao YL, Li SJ, Hao YQ, Yang R, Prima F. Corrosion behavior of biomedical Ti-24Nb-4Zr-8Sn alloy in different simulated body solutions. Mater Sci
Eng C 2013;33:2159–67. doi: http://dx.doi.org/10.1016/j.msec.2013.01.036.
[76] Sakaguchi N, Niinomi M, Akahori T, Takeda J, Toda H. Effect of Ta content on mechanical properties of Ti-30Nb-XTa-5Zr. Mater Sci Eng C
2005;25:370–6. doi: http://dx.doi.org/10.1016/j.msec.2005.04.003.
[77] Jiang A, Yohannan A, Nnolim NO, Tyson TA, Axe L, Lee SL. Investigation of the structure of b-tantalum. Thin Solid Films 2003;437:116–22.
[78] Knepper R. Thermomechanical behavior and microstructure evolution of Tantalum thin films during the beta-alpha phase transformation. Cornell
University; 2007.
[79] Zardiackas LD, Parsell DE, Dillon LD, Mitchell DW, Nunnery LA, Poggie R. Structure, metallurgy, and mechanical properties of a porous tantalum foam.
J Biomed Mater Res 2001;58:180–7.
[80] Hanzlik JA, Day JS. Bone ingrowth in well-fixed retrieved porous tantalum implants. J Arthroplasty 2013;28:922–7. doi: http://dx.doi.org/10.1016/j.
arth.2013.01.035.
[81] Findlay DM, Welldon K, Atkins GJ, Howie DW, Zannettino ACW, Bobyn D. The proliferation and phenotypic expression of human osteoblasts on
tantalum metal. Biomaterials 2004;25:2215–27. doi: http://dx.doi.org/10.1016/j.biomaterials.2003.09.005.
[82] Balagna C, Faga MG, Spriano S. Tantalum-based multilayer coating on cobalt alloys in total hip and knee replacement. Mater Sci Eng C
2012;32:887–95. doi: http://dx.doi.org/10.1016/j.msec.2012.02.007.
[83] Spriano S BS, Spriano S, Bugliosi S. Medical prosthetic devices presenting enhanced biocompatibility and wear resistance, based on cobalt alloys and
process for their preparation; 2006.
[84] Matsuno H, Yokoyama A, Watari F, Uo M, Kawasaki T. Biocompatibility and osteogenesis of refractory metal implants, titanium, hafnium, niobium,
tantalum and rhenium. Biomaterials 2001;22:1253–62. doi: http://dx.doi.org/10.1016/S0142-9612(00)00275-1.
[85] Li C, Zhan Y, Jiang W. Zr–Si biomaterials with high strength and low elastic modulus. Mater Des 2011;32:4598–602. doi: http://dx.doi.org/10.1016/
j.matdes.2011.03.072.
[86] Kondo R, Nomura N, Suyalatu, Tsutsumi Y, Doi H, Hanawa T. Microstructure and mechanical properties of as-cast Zr-Nb alloys. Acta Biomater
2011;7:4278–84. doi: http://dx.doi.org/10.1016/j.actbio.2011.07.020.
[87] Nie L, Zhan Y, Liu H, Tang C. Novel b-type Zr–Mo–Ti alloys for biological hard tissue replacements. Mater Des 2014;53:8–12. doi: http://dx.doi.org/
10.1016/j.matdes.2013.07.008.
[88] Hua N, Huang L, Wang J, Cao Y, He W, Pang S, et al. Corrosion behavior and in vitro biocompatibility of Zr–Al–Co–Ag bulk metallic glasses: an
experimental case study. J Non Cryst Solids 2012;358:1599–604. doi: http://dx.doi.org/10.1016/j.jnoncrysol.2012.04.022.
[89] Inoue A, Kawase D, Tsai A, Zhang T, Masumoto T. Stability and transformation to crystalline phases of amorphous Zr-Al-Cu alloys with significant
supercooled liquid region. Mater Sci Eng A 1994;178:255–63. doi: http://dx.doi.org/10.1016/0921-5093(94)90551-7.
[90] Lu XY, Huang L, Pang SJ, Zhang T. Formation and biocorrosion behavior of Zr-Al-Co-Nb bulk metallic glasses. Chinese Sci Bull 2012;57:1723–7. doi:
http://dx.doi.org/10.1007/s11434-012-5027-0.
[91] Bachtar F, Chen X, Hisada T. Finite element contact analysis of the hip joint. Med Biol Eng Comput 2006;44:643–51. doi: http://dx.doi.org/10.1007/
s11517-006-0074-9.
[92] Jilani A, Bendjaballah MZ. Biomechanics of human tibio-femoral joint in axial rotation. Knee 1997;4:203–13.
[93] Kutzner I, Heinlein B, Graichen F, Bender A, Rohlmann A, Halder A, et al. Loading of the knee joint during activities of daily living measured in vivo in
five subjects. J Biomech 2010;43:2164–73. doi: http://dx.doi.org/10.1016/j.jbiomech.2010.03.04.
[94] Carr BC, Goswami T. Knee implants – review of models and biomechanics. Mater Des 2009;30:398–413. doi: http://dx.doi.org/10.1016/
j.matdes.2008.03.032.
L. Kunčická et al. / Progress in Materials Science 88 (2017) 232–280 273

[95] Staiger MP, Pietak AM, Huadmai J, Dias G. Magnesium and its alloys as orthopedic biomaterials: a review. Biomaterials 2006;27:1728–34. doi: http://
dx.doi.org/10.1016/j.biomaterials.2005.10.003.
[96] Gu XN, Zheng YF. A review on magnesium alloys as biodegradable materials. Front Mater Sci China 2010;4:111–5. doi: http://dx.doi.org/10.1007/
s11706-010-0024-1.
[97] Hao L, Harris R. Customised implants for bone replacement and growth. In: Bártolo PJ, Bidanda B, editors. Bio-Materials Prototyp Appl Med.. New
York: Springer; 2008.
[98] Yang K, Ren Y. Nickel-free austenitic stainless steels for medical applications. Sci Technol Adv Mater 2010;11:14105. doi: http://dx.doi.org/10.1088/
1468-6996/11/1/014105.
[99] Kunčická L, Kocich R, Drápala J, Andreyachshenko VA. FEM simulations and comparison of the ecap and ECAP-PBP influence on Ti6Al4V alloy’s
deformation behaviour. In: Met 2013 22nd int met mater conf. p. 391–6.
[100] Kocich R, Kursa M, Szurman I, Dlouhý A. The influence of imposed strain on the development of microstructure and transformation characteristics of
Ni–Ti shape memory alloys. J Alloys Compd 2011;509:2716–22. doi: http://dx.doi.org/10.1016/j.jallcom.2010.12.003.
[101] Gabriel SB, Panaino JVP, Santos ID, Araujo LS, Mei PR, De Almeida LH, et al. Characterization of a new beta titanium alloy, Ti-12Mo-3Nb, for biomedical
applications. J Alloys Compd 2012;536:S208–10. doi: http://dx.doi.org/10.1016/j.jallcom.2011.11.035.
[102] Dai SJ, Wang Y, Chen F, Yu XQ, Zhang YF. Influence of Zr content on microstructure and mechanical properties of implant Ti-35Nb-4Sn-6Mo-xZr
alloys. Trans Nonferrous Met Soc China (English Ed) 2013;23:1299–303. doi: http://dx.doi.org/10.1016/S1003-6326(13)62597-2.
[103] Ramarolahy a, Castany P, Prima F, Laheurte P, Péron I, Gloriant T. Microstructure and mechanical behavior of superelastic Ti-24Nb-0.5O and Ti-24Nb-
0.5N biomedical alloys. J Mech Behav Biomed Mater 2012;9:83–90. doi: http://dx.doi.org/10.1016/j.jmbbm.2012.01.017.
[104] Papakyriacou M, Mayer H, Plenk H, Stanzl-Tschegg S. Cyclic plastic deformation of tantalum and niobium at very high numbers of cycles. Mater Sci
Eng A 2002;325:520–4. doi: http://dx.doi.org/10.1016/S0921-5093(01)01446-0.
[105] Papakyriacou M, Mayer H, Pypen C. Effects of surface treatments on high cycle corrosion fatigue of metallic implant materials. Int J Fatigue
2000;22:873–86.
[106] Nie L, Zhan Y, Hu T, Chen X, Wang C. Β-Type Zr-Nb-Ti biomedical materials with high plasticity and low modulus for hard tissue replacements. J Mech
Behav Biomed Mater 2014;29:1–6. doi: http://dx.doi.org/10.1016/j.jmbbm.2013.08.019.
[107] Suyalatu, Kondo R, Tsutsumi Y, Doi H, Nomura N, Hanawa T. Effects of phase constitution on magnetic susceptibility and mechanical properties of Zr-
rich Zr-Mo alloys. Acta Biomater 2011;7:4259–66. doi: http://dx.doi.org/10.1016/j.actbio.2011.07.005.
[108] Ng HP, Haase C, Lapovok R, Estrin Y. Improving sinterability of Ti–6Al–4V from blended elemental powders through equal channel angular pressing.
Mater Sci Eng A 2013;565:396–404. doi: http://dx.doi.org/10.1016/j.msea.2012.12.071.
[109] Haase C, Lapovok R, Ng HP, Estrin Y. Production of Ti–6Al–4V billet through compaction of blended elemental powders by equal-channel angular
pressing. Mater Sci Eng A 2012;550:263–72. doi: http://dx.doi.org/10.1016/j.msea.2012.04.068.
[110] Zhou YL, Niinomi M, Akahori T. Changes in mechanical properties of Ti alloys in relation to alloying additions of Ta and Hf. Mater Sci Eng A 2008;483–
484:153–6. doi: http://dx.doi.org/10.1016/j.msea.2006.09.173.
[111] Kim H, Ra T, Yeo I, Bang H. Microstructure, elastic modulus and tensile properties of Ti-Nb-O alloy system. J Mater Sci Technol 2008;24:33–6.
[112] Javanbakht M, Hadianfard MJ, Salahinejad E. Microstructure and mechanical properties of a new group of nanocrystalline medical-grade stainless
steels prepared by powder metallurgy. J Alloys Compd 2015;624:17–21. doi: http://dx.doi.org/10.1016/j.jallcom.2014.11.080.
[113] Shankar P, Palanichamy P, Jayakumar T, Raj B, Ranganathan S. Nitrogen redistribution, microstructure, and elastic constant evaluation using
ultrasonics in aged 316LN stainless steels. Metall Mater Trans A 2001;32:2959–68. doi: http://dx.doi.org/10.1007/s11661-001-0170-2.
[114] Salahinejad E, Amini R, Hadianfard MJ. Contribution of nitrogen concentration to compressive elastic modulus of 18Cr-12Mn-xN austenitic stainless
steels developed by powder metallurgy. Mater Des 2010;31:2241–4. doi: http://dx.doi.org/10.1016/j.matdes.2009.10.016.
[115] Bender S, Chalivendra V, Rahbar N, El Wakil S. Mechanical characterization and modeling of graded porous stainless steel specimens for possible bone
implant applications. Int J Eng Sci 2012;53:67–73. doi: http://dx.doi.org/10.1016/j.ijengsci.2012.01.004.
[116] Henriques B, Gasik M, Souza JCM, Nascimento RM, Soares D, Silva FS. Mechanical and thermal properties of hot pressed CoCrMo-porcelain composites
developed for prosthetic dentistry. J Mech Behav Biomed Mater 2014;30:103–10. doi: http://dx.doi.org/10.1016/j.jmbbm.2013.10.023.
[117] He G, Hagiwara M. Bimodal structured Ti-base alloy with large elasticity and low Young’s modulus. Mater Sci Eng C 2005;25:290–5. doi: http://dx.
doi.org/10.1016/j.msec.2005.03.001.
[118] Reig L, Amigó V, Busquets DJ, Calero Ja. Development of porous Ti6Al4V samples by microsphere sintering. J Mater Process Technol 2012;212:3–7.
doi: http://dx.doi.org/10.1016/j.jmatprotec.2011.06.026.
[119] Niinomi M. Mechanical biocompatibilities of titanium alloys for biomedical applications. J Mech Behav Biomed Mater 2008;1:30–42. doi: http://dx.
doi.org/10.1016/j.jmbbm.2007.07.001.
[120] Ozaki T, Matsumoto H, Watanabe S, Hanada S. Beta Ti alloys with low young’s modulus. Mater Trans 2004;45:2776–9.
[121] Nouri A, Hodgson PD, Wen CE. Effect of process control agent on the porous structure and mechanical properties of a biomedical Ti-Sn-Nb alloy
produced by powder metallurgy. Acta Biomater 2010;6:1630–9. doi: http://dx.doi.org/10.1016/j.actbio.2009.10.005.
[122] Zhou YL, Niinomi M. Ti-25Ta alloy with the best mechanical compatibility in Ti-Ta alloys for biomedical applications. Mater Sci Eng C
2009;29:1061–5. doi: http://dx.doi.org/10.1016/j.msec.2008.09.012.
[123] Zhou YL, Niinomi M. Microstructures and mechanical properties of Ti-50 mass% Ta alloy for biomedical applications. J Alloys Compd
2008;466:535–42. doi: http://dx.doi.org/10.1016/j.jallcom.2007.11.090.
[124] Balla VK, Bodhak S, Bose S, Bandyopadhyay A. Porous tantalum structures for bone implants: fabrication, mechanical and in vitro biological
properties. Acta Biomater 2010;6:3349–59. doi: http://dx.doi.org/10.1016/j.actbio.2010.01.046.
[125] Wauthle R, van der Stok J, Amin Yavari S, Van Humbeeck J, Kruth J-P, Zadpoor AA, et al. Additively manufactured porous tantalum implants. Acta
Biomater 2014;14:217–25. doi: http://dx.doi.org/10.1016/j.actbio.2014.12.003.
[126] Fischer A, Weiss S, Wimmer Ma. The tribological difference between biomedical steels and CoCrMo-alloys. J Mech Behav Biomed Mater
2012;9:50–62. doi: http://dx.doi.org/10.1016/j.jmbbm.2012.01.007.
[127] Younesi M, Bahrololoom ME. Optimizations of wear resistance and toughness of hydroxyapatite nickel free stainless steel new bio-composites for
using in total joint replacement. Mater Des 2010;31:234–43. doi: http://dx.doi.org/10.1016/j.matdes.2009.06.037.
[128] Kim Y-H. Comparison of polyethylene wear associated with cobalt-chromium and zirconia heads after total hip replacement. A prospective,
randomized study. J Bone Joint Surg Am 2005;87:1769–76. doi: http://dx.doi.org/10.2106/JBJS.D.02572.
[129] Wang A, Polineni VK, Stark C, Dumbleton JH. Effect of femoral head surface roughness on the wear of ultrahigh molecular weight polyethylene
acetabular cups. J Arthroplasty 1998;13:615–20. doi: http://dx.doi.org/10.1016/S0883-5403(98)80002-8.
[130] Gremillard L, Martin L, Zych L, Crosnier E, Chevalier J, Charbouillot A, et al. Combining ageing and wear to assess the durability of zirconia-based
ceramic heads for total hip arthroplasty. Acta Biomater 2013;9:7545–55. doi: http://dx.doi.org/10.1016/j.actbio.2013.03.030.
[131] Figueiredo-Pina CGG, Dearnley Pa, Fisher J. UHMWPE wear response to apposing nitrogen S-phase coated and uncoated orthopaedic implant grade
stainless steel. Wear 2009;267:743–52. doi: http://dx.doi.org/10.1016/j.wear.2009.01.036.
[132] Thomann UI, Uggowitzer PJ. Wear–corrosion behavior of biocompatible austenitic stainless steels. Wear 2000;239:48–58. doi: http://dx.doi.org/
10.1016/S0043-1648(99)00372-5.
[133] Mallia B, Stüber M, Dearnley Pa. Character and chemical-wear response of high alloy austenitic stainless steel (Ortron 90) surface engineered with
magnetron sputtered Cr-B-N ternary alloy coatings. Thin Solid Films 2013;549:216–23. doi: http://dx.doi.org/10.1016/j.tsf.2013.08.093.
[134] McGrory BJ, Ruterbories JM, Pawar VD, Thomas RK, Salehi AB. Comparison of surface characteristics of retrieved cobalt-chromium femoral heads with
and without ion implantation. J Arthroplasty 2012;27:109–15. doi: http://dx.doi.org/10.1016/j.arth.2011.03.009.
[135] Ratner BD, Hoffman AS, Schoen FJ, Lemons JE. Biomaterials science: an introduction to materials in medicine. Academic Press; 2004.
274 L. Kunčická et al. / Progress in Materials Science 88 (2017) 232–280

[136] Saldívar-García AJ, López HF. Microstructural effects on the wear resistance of wrought and as-cast Co-Cr-Mo-C implant alloys. J Biomed Mater Res A
2005;74:269–74. doi: http://dx.doi.org/10.1002/jbm.a.30392.
[137] Hsu HC, Lian SS. Wear properties of Co-Cr-Mo-N plasma-melted surgical implant alloys. J Mater Process Technol 2003;138:231–5. doi: http://dx.doi.
org/10.1016/S0924-0136(03)00077-3.
[138] Doni Z, Alves aC, Toptan F, Gomes JR, Ramalho A, Buciumeanu M, et al. Dry sliding and tribocorrosion behaviour of hot pressed CoCrMo biomedical
alloy as compared with the cast CoCrMo and Ti6Al4V alloys. Mater Des 2013;52:47–57. doi: http://dx.doi.org/10.1016/j.matdes.2013.05.03.
[139] Cvijovic-Alagic I, Cvijovic Z, Mitrovic S, Panic V, Rakin M. Wear and corrosion behaviour of Ti-13Nb-13Zr and Ti-6Al-4V alloys in simulated
physiological solution. Corros Sci 2011;53:796–808. doi: http://dx.doi.org/10.1016/j.corsci.2010.11.014.
[140] Lee Y-S, Niinomi M, Nakai M, Narita K, Cho K. Predominant factor determining wear properties of b-type and (a+b)-type titanium alloys in metal-to-
metal contact for biomedical applications. J Mech Behav Biomed Mater 2015;41:208–20. doi: http://dx.doi.org/10.1016/j.jmbbm.2014.10.005.
[141] Tkachenko S, Datskevich O, Kulak L, Jacobson S, Engqvist H, Persson C. Wear and friction properties of experimental Ti–Si–Zr alloys for biomedical
applications. J Mech Behav Biomed Mater 2014;39:61–72. doi: http://dx.doi.org/10.1016/j.jmbbm.2014.07.011.
[142] Zhou YL, Niinomi M, Akahori T, Fukui H, Toda H. Corrosion resistance and biocompatibility of Ti-Ta alloys for biomedical applications. Mater Sci Eng A
2005;398:28–36. doi: http://dx.doi.org/10.1016/j.msea.2005.03.032.
[143] Attar H, Prashanth KG, Chaubey AK, Calin M, Zhang LC, Scudino S, et al. Comparison of wear properties of commercially pure titanium prepared by
selective laser melting and casting processes. Mater Lett 2015;142:38–41. doi: http://dx.doi.org/10.1016/j.matlet.2014.11.156.
[144] Güleryüz H, Çimenoǧlu H. Effect of thermal oxidation on corrosion and corrosion-wear behaviour of a Ti-6Al-4V alloy. Biomaterials
2004;25:3325–33. doi: http://dx.doi.org/10.1016/j.biomaterials.2003.10.009.
[145] Dearnley Pa, Dahm KL, Çimenoǧlu H. The corrosion-wear behaviour of thermally oxidised CP-Ti and Ti-6Al-4V. Wear 2004;256:469–79. doi: http://
dx.doi.org/10.1016/S0043-1648(03)00557-X.
[146] Dong H, Bell T. Enhanced wear resistance of titanium surfaces by a new thermal oxidation treatment. Wear 2000;238:131–7. doi: http://dx.doi.org/
10.1016/S0043-1648(99)00359-2.
[147] Mohan L, Anandan C. Wear and corrosion behavior of oxygen implanted biomedical titanium alloy Ti-13Nb-13Zr. Appl Surf Sci 2013;282:281–90.
doi: http://dx.doi.org/10.1016/j.apsusc.2013.05.120.
[148] Dittrick S, Balla VK, Bose S, Bandyopadhyay A. Wear performance of laser processed tantalum coatings. Mater Sci Eng C 2011;31:1832–5. doi: http://
dx.doi.org/10.1016/j.msec.2011.08.017.
[149] Byeli AV, Kukareko VA, Kononov AG. Titanium and zirconium based alloys modified by intensive plastic deformation and nitrogen ion implantation
for biocompatible implants. J Mech Behav Biomed Mater 2012;6:89–94. doi: http://dx.doi.org/10.1016/j.jmbbm.2011.05.034.
[150] Pandey AK, Biswas K. Influence of sintering parameters on tribological properties of ceria stabilized zirconia bio-ceramics. Ceram Int
2011;37:257–64. doi: http://dx.doi.org/10.1016/j.ceramint.2010.08.041.
[151] Teoh S. Fatigue of biomaterials: a review. Int J Fatigue 2000;22:825–37. doi: http://dx.doi.org/10.1016/S0142-1123(00)00052-9.
[152] Stráský J, Janeček M, Harcuba P, Bukovina M, Wagner L. The effect of microstructure on fatigue performance of Ti-6Al-4V alloy after EDM surface
treatment for application in orthopaedics. J Mech Behav Biomed Mater 2011;4:1955–62. doi: http://dx.doi.org/10.1016/j.jmbbm.2011.06.012.
[153] Lin CW, Ju CP, Chern Lin JH. A comparison of the fatigue behavior of cast Ti-7.5Mo with c.p. titanium, Ti-6Al-4V and Ti-13Nb-13Zr alloys. Biomaterials
2005;26:2899–907. doi: http://dx.doi.org/10.1016/j.biomaterials.2004.09.007.
[154] Chevalier J, Taddei P, Gremillard L, Deville S, Fantozzi G, Bartolomé JF, et al. Reliability assessment in advanced nanocomposite materials for
orthopaedic applications. J Mech Behav Biomed Mater 2011;4:303–14. doi: http://dx.doi.org/10.1016/j.jmbbm.2010.10.010.
[155] Lin X, Haicheng G. High cycle fatigue properties and microstructure of zirconium and zircaloy-4 under reversal bending. Mater Sci Eng A
1998;252:166–73. doi: http://dx.doi.org/10.1016/S0921-5093(98)00688-1.
[156] Nikulin SA, Markelov VA, Gusev AY, Nechaykina TA, Rozhnov aB, Rogachev SO, et al. Low-cycle fatigue tests of zirconium alloys using a dynamic
mechanical analyzer. Int J Fatigue 2013;48:187–91. doi: http://dx.doi.org/10.1016/j.ijfatigue.2012.10.019.
[157] Ryu JJ, Shrotriya P. Influence of roughness on surface instability of medical grade cobalt-chromium alloy (CoCrMo) during contact corrosion-fatigue.
Appl Surf Sci 2013;273:536–41. doi: http://dx.doi.org/10.1016/j.apsusc.2013.02.076.
[158] Noli F, Misaelides P, Lagoyannis a, Pichon L, Ozturk O. Use of combination of accelerator-based ion-beam analysis techniques to the investigation of
the corrosion behavior of CoCrMo alloy. Nucl Instrum Meth Phys Res Sect B Beam Interact Mater Atoms 2014;331:125–9. doi: http://dx.doi.org/
10.1016/j.nimb.2013.11.039.
[159] Høl PJ, Mølster A, Gjerdet NR. Should the galvanic combination of titanium and stainless steel surgical implants be avoided? Injury 2008;39:161–9.
doi: http://dx.doi.org/10.1016/j.injury.2007.07.015.
[160] Metikoš-Huković M, Babić R. Passivation and corrosion behaviours of cobalt and cobalt–chromium–molybdenum alloy. Corros Sci 2007;49:3570–9.
doi: http://dx.doi.org/10.1016/j.corsci.2007.03.023.
[161] Espallargas N, Torres C, Muñoz AI. A metal ion release study of CoCrMo exposed to corrosion and tribocorrosion conditions in simulated body fluids.
Wear 2014:1–10. doi: http://dx.doi.org/10.1016/j.wear.2014.12.030.
[162] Dorn U, Neumann D, Frank M. Corrosion behavior of tantalum-coated cobalt-chromium modular necks compared to titanium modular necks in a
simulator test. J Arthroplasty 2014;29:831–5. doi: http://dx.doi.org/10.1016/j.arth.2013.08.022.
[163] Peng DQ, Bai XD, Chen BS. Surface analysis and corrosion behavior of zirconium samples implanted with yttrium and lanthanum. Surf Coat Technol
2005;190:440–7. doi: http://dx.doi.org/10.1016/j.surfcoat.2004.05.024.
[164] Holzapfel BM, Reichert JC, Schantz JT, Gbureck U, Rackwitz L, N??th U, et al. How smart do biomaterials need to be? A translational science and
clinical point of view. Adv Drug Deliv Rev 2013;65:581–603. doi: http://dx.doi.org/10.1016/j.addr.2012.07.009.
[165] Chen Q, Liang S, Thouas GA. Elastomeric biomaterials for tissue engineering. Prog Polym Sci 2013;38:584–671. doi: http://dx.doi.org/10.1016/j.
progpolymsci.2012.05.003.
[166] Nair LS, Laurencin CT. Biodegradable polymers as biomaterials. Prog Polym Sci 2007;32:762–98. doi: http://dx.doi.org/10.1016/j.
progpolymsci.2007.05.017.
[167] McKellop H, Shen FW, Lu B, Campbell P, Salovey R. Development of an extremely wear-resistant ultra high molecular weight polyethylene for total
hip replacements. J Orthop Res 1999;17:157–67. doi: http://dx.doi.org/10.1002/jor.1100170203.
[168] Heyse TJ, Chen DX, Kelly N, Boettner F, Wright TM, Haas SB. Matched-pair total knee arthroplasty retrieval analysis: Oxidized zirconium vs. CoCrMo.
Knee 2011;18:448–52. doi: http://dx.doi.org/10.1016/j.knee.2010.08.011.
[169] Sutha S, Karunakaran G, Rajendran V. Enhancement of antimicrobial and long-term biostability of the zinc-incorporated hydroxyapatite coated 316L
stainless steel implant for biomedical application. Ceram Int 2013;39:5205–12. doi: http://dx.doi.org/10.1016/j.ceramint.2012.12.019.
[170] Ajeesh M, Francis BF, Annie J, Harikrishna Varma PR. Nano iron oxide-hydroxyapatite composite ceramics with enhanced radiopacity. J Mater Sci
Mater Med 2010;21:1427–34. doi: http://dx.doi.org/10.1007/s10856-010-4005-9.
[171] Chen Q, Zhu C, Thouas Ga. Progress and challenges in biomaterials used for bone tissue engineering: bioactive glasses and elastomeric composites.
Prog Biomater 2012;1:2. doi: http://dx.doi.org/10.1186/2194-0517-1-2.
[172] Li Z, Kawashita M. Current progress in inorganic artificial biomaterials. J Artif Organs 2011;14:163–70. doi: http://dx.doi.org/10.1007/s10047-011-
0585-5.
[173] Biesiekierski A, Wang J, Gepreel MA-H, Wen C. A new look at biomedical Ti-based shape memory alloys. Acta Biomater 2012;8:1661–9.
[174] van Hove RP, Sierevelt IN, van Royen BJ, Nolte PA. Titanium-nitride coating of orthopaedic implants: a review of the literature. Biomed Res Int
2015;2015:485975. doi: http://dx.doi.org/10.1155/2015/485975.
[175] Gispert MP, Serro AP, Colaço R, Pires E, Saramago B. The effect of roughness on the tribological behavior of the prosthetic pair UHMWPE/TiN-coated
stainless steel. J Biomed Mater Res B Appl Biomater 2008;84:98–107. doi: http://dx.doi.org/10.1002/jbm.b.30849.
L. Kunčická et al. / Progress in Materials Science 88 (2017) 232–280 275

[176] Yildiz F, Yetim aF, Alsaran a, Çelik a. Plasma nitriding behavior of Ti6Al4V orthopedic alloy. Surf Coat Technol 2008;202:2471–6. doi: http://dx.doi.
org/10.1016/j.surfcoat.2007.08.004.
[177] Yildiz F, Yetim aF, Alsaran a, Efeoglu I. Wear and corrosion behaviour of various surface treated medical grade titanium alloy in bio-simulated
environment. Wear 2009;267:695–701. doi: http://dx.doi.org/10.1016/j.wear.2009.01.056.
[178] Werner Z, Barlak M, Gra˛dzka-Dahlke M, Diduszko R, Szymczyk W, Da˛browski J, et al. The effect of ion implantation on the wear of Co–Cr–Mo alloy.
Vacuum 2007;81:1191–4. doi: http://dx.doi.org/10.1016/j.vacuum.2007.01.014.
[179] Biomedical N. Technologies - N2 Biomedical N2 Biomedical 2016. <http://n2bio.com/our-technologies/> [accessed February 20, 2016].
[180] Gu K, Wang J, Zhou Y. Effect of cryogenic treatment on wear resistance of Ti-6Al-4V alloy for biomedical applications. J Mech Behav Biomed Mater
2014;30:131–9. doi: http://dx.doi.org/10.1016/j.jmbbm.2013.11.003.
[181] Board NE. 12th Annual Report 2015 National Joint Registry for England, Wales, Northern Ireland and the Isle of Man; 2015.
[182] Pezzotti G, Yamamoto K. Artificial hip joints: the biomaterials challenge. J Mech Behav Biomed Mater 2014;31:3–20. doi: http://dx.doi.org/10.1016/j.
jmbbm.2013.06.001.
[183] Arola D, Alade AE, Weber W. Improving fatigue strength of metals using abrasive waterjet peening. Mach Sci Technol 2006;10:197–218. doi: http://
dx.doi.org/10.1080/10910340600710105.
[184] Sonntag R, Reinders J, Gibmeier J, Kretzer JP. Fatigue performance of medical Ti6Al4V alloy after mechanical surface treatments. PLoS ONE 2015;10:
e0121963. doi: http://dx.doi.org/10.1371/journal.pone.0121963.
[185] Ye X, Ye Y, Tang G. Effect of electropulsing treatment and ultrasonic striking treatment on the mechanical properties and microstructure of
biomedical ti-6Al-4V alloy. J Mech Behav Biomed Mater 2014;40:287–96. doi: http://dx.doi.org/10.1016/j.jmbbm.2014.08.022.
[186] Sioshansi P, Tobin EJ. Surface treatment of biomaterials by ion beam processes. Surf Coat Technol 1996;83:175–82. doi: http://dx.doi.org/10.1016/
0257-8972(95)02838-2.
[187] Strasky J, Havlikova J, Bacakova L, Harcuba P, Mhaede M, Janecek M. Characterization of electric discharge machining, subsequent etching and shot-
peening as a surface treatment for orthopedic implants. Appl Surf Sci 2013;281:73–8. doi: http://dx.doi.org/10.1016/j.apsusc.2013.02.053.
[188] Ponche A, Bigerelle M, Anselme K. Relative influence of surface topography and surface chemistry on cell response to bone implant materials. Part 1:
Physico-chemical effects. Proc Inst Mech Eng Part H-J Eng Med 2010;224:1471–86. doi: http://dx.doi.org/10.1243/09544119jeim900.
[189] Rani VVD, Vinoth-Kumar L, Anitha VC, Manzoor K, Deepthy M, Shantikumar VN. Osteointegration of titanium implant is sensitive to specific
nanostructure morphology. Acta Biomater 2012;8:1976–89. doi: http://dx.doi.org/10.1016/j.actbio.2012.01.021.
[190] Song Y, Beaupre G, Goodman S. Osseointegration of total hip arthroplasties: studies in humans and animals. J Long-Term Eff 1998.
[191] Mavrogenis A, Dimitriou R. Biology of implant osseointegration. J Musculoskelet Neuronal 2009.
[192] Karageorgiou V, Kaplan D. Porosity of 3D biomaterial scaffolds and osteogenesis. Biomaterials 2005;26:5474–91. doi: http://dx.doi.org/10.1016/j.
biomaterials.2005.02.002.
[193] Otsuki B, Takemoto M, Fujibayashi S, Neo M, Kokubo T, Nakamura T. Pore throat size and connectivity determine bone and tissue ingrowth into
porous implants: three-dimensional micro-CT based structural analyses of porous bioactive titanium implants. Biomaterials 2006;27:5892–900. doi:
http://dx.doi.org/10.1016/j.biomaterials.2006.08.013.
[194] Taniguchi N, Fujibayashi S, Takemoto M, Sasaki K, Otsuki B, Nakamura T, et al. Effect of pore size on bone ingrowth into porous titanium implants
fabricated by additive manufacturing: an in vivo experiment. Mater Sci Eng C Mater Biol Appl 2016;59:690–701. doi: http://dx.doi.org/10.1016/j.
msec.2015.10.069.
[195] Ferraris S, Spriano S, Pan G, Venturello A, Bianchi CL, Chiesa R, et al. Surface modification of Ti-6Al-4V alloy for biomineralization and specific
biological response: Part I, inorganic modification. J Mater Sci Med 2011;22:533–45. doi: http://dx.doi.org/10.1007/s10856-011-4246-2.
[196] Patil N, Lee K, Goodman S. Porous tantalum in hip and knee reconstructive surgery. J Biomed Mater 2009.
[197] Basalah A, Shanjani Y, Esmaeili S, Toyserkani E. Characterizations of additive manufactured porous titanium implants. J Biomed Mater Res Part B-Appl
Biomater 2012;100B:1970–9. doi: http://dx.doi.org/10.1002/jbm.b.32764.
[198] Hacking S, Bobyn J, Toh K. Fibrous tissue ingrowth and attachment to porous tantalum. J Biomed 2000.
[199] Bi H, Yu C, Cao P, He YH. Porous Ti-6Al-4V alloy prepared by a press-and-sinter process. In: Qian M, editor. Powder Metall Titan Powder Process
Consol Metall Titan, vol. 520. p. 76–81. doi: http://dx.doi.org/10.4028/www.scientific.net/KEM.520.76.
[200] Turner T, Sumner D, Urban R. A comparative study of porous coatings in a weight-bearing total hip-arthroplasty model. J Bone 1986.
[201] Engh C. Porous-coated hip replacement. The factors governing bone ingrowth, stress shielding, and clinical results. J Bone 1987.
[202] Lefebvre L-P, Baril E. Properties of titanium foams for biomedical applications. Adv Eng Mater 2013;15:159–65. doi: http://dx.doi.org/10.1002/
adem.201200154.
[203] Chu PK. Plasma-treated biomaterials. IEEE Trans Plasma Sci 2007;35:181–7. doi: http://dx.doi.org/10.1109/tps.2006.888587.
[204] Chu P. Plasma-surface modification of biomaterials. Mater Sci Eng R Reports 2002;36:143–206. doi: http://dx.doi.org/10.1016/S0927-796X(02)
00004-9.
[205] Emerson RH, Sanders SB, Head WC, Higgins L. Effect of circumferential plasma-spray porous coating on the rate of femoral osteolysis after total hip
arthroplasty. J Bone Joint Surg Am 1999;81:1291–8.
[206] Wang M. Surface modification of metallic biomaterials for orthopaedic applications. Mater Sci 2009;618–619:285–90. doi: http://dx.doi.org/10.4028/
www.scientific.net/MSF.618-619.285.
[207] Wang M. Surface modification of biomaterials and tissue engineering scaffolds for enhanced osteoconductivity. IFMBE Proc 2007. doi: http://dx.doi.
org/10.1007/978-3-540-68017-8.
[208] Stryker. Revision Knee Systems - Global Modular Replacement System 2016. <http://www.stryker.com/en-us/products/Orthopaedics/
KneeReplacement/Revision/GlobalModularReplacementSystem/index.htm> [accessed December 15, 2016].
[209] Won Y-Y, Dorr LD, Wan Z. Comparison of proximal porous-coated and grit-blasted surfaces of hydroxyapatite-coated stems. J Bone Joint Surg Am
2004;86–A:124–8.
[210] Bolognesi MP, Pietrobon R, Clifford PE, Vail TP. Comparison of a hydroxyapatite-coated sleeve and a porous-coated sleeve with a modular revision hip
stem. A prospective, randomized study. J Bone Joint Surg Am 2004;86–A:2720–5.
[211] Park Y-S, Lim S-J. Long-term comparison of porous and hydroxyapatite sleeves in femoral revision using the S-ROM modular stem. Hip Int
2010;20:179–86. doi: http://dx.doi.org/10.5301/HIP.2010.1544.
[212] Aparicio C, Gil FJ, Planell JA, Engel E. Human-osteoblast proliferation and differentiation on grit-blasted and bioactive titanium for dental applications.
J Mater Sci Mater Med n.d.; 13:1105–11. doi:http://dx.doi.org/10.1023/A:1021173500990.
[213] Bloyer DR, Gomez-Vega JM, Saiz E, Mcnaney JM, Cannon RM, Tomsia AP. Fabrication and characterization of a bioactive glass coating on titanium
implant alloys; 1999.
[214] Sharifi E, Azami M, Kajbafzadeh A-M, Moztarzadeh F, Faridi-Majidi R, Shamousi A, et al. Preparation of a biomimetic composite scaffold from
gelatin/collagen and bioactive glass fibers for bone tissue engineering. Mater Sci Eng C 2016. doi: http://dx.doi.org/10.1016/j.msec.2015.09.037.
[215] Chen Q, Yang Y, Pérez de Larraya U, Garmendia N, Virtanen S, Boccaccini AR. Electrophoretic co-deposition of cellulose nanocrystals-45S5 bioactive
glass nanocomposite coatings on stainless steel. Appl Surf Sci 2016. doi: http://dx.doi.org/10.1016/j.apsusc.2015.11.160.
[216] Blaker JJ, Nazhat SN, Boccaccini AR. Development and characterisation of silver-doped bioactive glass-coated sutures for tissue engineering and
wound healing applications. Biomaterials 2004. doi: http://dx.doi.org/10.1016/j.biomaterials.2003.08.007.
[217] Kiefer K, Amlung M, Aktas OC, de Oliveira PW, Abdul-Khaliq H. Novel glass-like coatings for cardiovascular implant application: preparation,
characterization and cellular interaction. Mater Sci Eng C 2016. doi: http://dx.doi.org/10.1016/j.msec.2015.09.063.
[218] Vitale-Brovarone C, Baino F, Tallia F, Gervasio C, Verné E. Bioactive glass-derived trabecular coating: a smart solution for enhancing osteointegration
of prosthetic elements n.d. doi:http://dx.doi.org/10.1007/s10856-012-4643-1.
276 L. Kunčická et al. / Progress in Materials Science 88 (2017) 232–280

[219] del Val J, López-Cancelos R, Riveiro A, Badaoui A, Lusquiños F, Quintero F, et al. On the fabrication of bioactive glass implants for bone regeneration by
laser assisted rapid prototyping based on laser cladding. Ceram Int 2016. doi: http://dx.doi.org/10.1016/j.ceramint.2015.10.009.
[220] Al-Noaman A, Rawlinson SCF, Hill RG. The role of MgO on thermal properties, structure and bioactivity of bioactive glass coating for dental implants. J
Non Cryst Solids 2012. doi: http://dx.doi.org/10.1016/j.jnoncrysol.2012.07.039.
[221] Vilardell AM, Cinca N, Concustell A, Dosta S, Cano IG, Guilemany JM. Cold spray as an emerging technology for biocompatible and antibacterial
coatings: state of art. J Mater Sci 2015;50:4441–62. doi: http://dx.doi.org/10.1007/s10853-015-9013-1.
[222] Hou N, Perinpanayagam H, Mozumder M, Zhu J. Novel development of biocompatible coatings for bone implants. Coatings 2015;5:737–57.
[223] Wang X, Xu S, Zhou S, Xu W, Leary M, Choong P, et al. Topological design and additive manufacturing of porous metals for bone scaffolds and
orthopaedic implants: a review. Biomaterials 2016;83:127–41.
[224] Abbasi Z, Bahrololoum ME, Bagheri R, Shariat MH. Characterization of the bioactive and mechanical behavior of dental ceramic/sol–gel derived
bioactive glass mixtures. J Mech Behav Biomed Mater 2016. doi: http://dx.doi.org/10.1016/j.jmbbm.2015.09.025.
[225] Sola A, Bellucci D, Cannillo V. Functionally graded materials for orthopedic applications – an update on design and manufacturing. Biotechnol Adv
2016.
[226] Kocich R, Macháčková A, Fojtík F. Comparison of strain and stress conditions in conventional and ARB rolling processes. Int J Mech Sci 2012;64:54–61.
doi: http://dx.doi.org/10.1016/j.ijmecsci.2012.08.003.
[227] Valiev RZ, Semenova IP, Jakushina E, Latysh VV, Rack H, Lowe TC, et al. Nanostructured SPD processed titanium for medical implants. Nanomater Sev
Plast Deform IV 2008;584–586:49–54.
[228] Ozaltin K, Chrominski W, Kulczyk M, Panigrahi A, Horky J, Zehetbauer M, et al. Enhancement of mechanical properties of biocompatible Ti-45Nb alloy
by hydrostatic extrusion. J Mater Sci 2014;49:6930–6. doi: http://dx.doi.org/10.1007/s10853-014-8397-7.
[229] Mazzoli A. Selective laser sintering in biomedical engineering. Med Biol Eng Comput 2013;51:245–56. doi: http://dx.doi.org/10.1007/s11517-012-
1001-x.
[230] Marshall GJ, Thompson SM, Shamsaei N. Data indicating temperature response of Ti–6Al–4V thin-walled structure during its additive manufacture
via Laser Engineered Net Shaping. Data Br 2016;7:697–703. doi: http://dx.doi.org/10.1016/j.dib.2016.02.084.
[231] Mok S-W, Nizak R, Fu S-C, Ho K-WK, Qin L, Saris DBF, et al. From the printer: potential of three-dimensional printing for orthopaedic applications. J
Orthop Transl 2016;6:42–9. doi: http://dx.doi.org/10.1016/j.jot.2016.04.003.
[232] Malik HH, Darwood ARJ, Shaunak S, Kulatilake P, El-Hilly AA, Mulki O, et al. Three-dimensional printing in surgery: a review of current surgical
applications. J Surg Res 2015;199:512–22. doi: http://dx.doi.org/10.1016/j.jss.2015.06.051.
[233] Kaushik P, Lehmann J, Nadif M. State of the art in control of inclusions, their characterization, and future requirements. Metall Mater Trans B Process
Metall Mater Process Sci 2012;43:710–25. doi: http://dx.doi.org/10.1007/s11663-012-9646-2.
[234] Yan P, Huang S, Pandelaers L, Dyck J, Guo M, Blanpain B. Effect of the CaO-Al2O3-based top slag on the cleanliness of stainless steel during secondary
metallurgy. Metall Mater Trans B Process Metall Mater Process Sci 2013;44:1105–19. doi: http://dx.doi.org/10.1007/s11663-013-9898-5.
[235] Grong O, Kolbeinsen L, van der Eijk C, Tranell G. Microstructure control of steels through dispersoid metallurgy using novel grain refining alloys. Isij
Int 2006;46:824–31. doi: http://dx.doi.org/10.2355/isijinternational.46.824.
[236] Wang C, Gao H, Dai Y, Ruan X, Wang J, Sun B. Grain refining of 409L ferritic stainless steel using Fe-Ti-N master alloy. Metall Mater Trans a-Physical
Metall Mater Sci 2010;41A:1616–20. doi: http://dx.doi.org/10.1007/s11661-010-0228-0.
[237] Balachandran G, Bhatia ML, Ballal NB, Rao PK. Processing nickel free high nitrogen austenitic stainless steels through conventional electroslag
remelting process. ISIJ Int 2000;40:478–83. doi: http://dx.doi.org/10.2355/isijinternational.40.478.
[238] Mengucci P, Barucca G, Gatto A, Bassoli E, Denti L, Fiori F, et al. Effects of thermal treatments on microstructure and mechanical properties of a Co–Cr–
Mo–W biomedical alloy produced by laser sintering. J Mech Behav Biomed Mater 2016;60:106–17.
[239] Elahinia MH, Hashemi M, Tabesh M, Bhaduri SB. Manufacturing and processing of NiTi implants: a review. Prog Mater Sci 2012;57:911–46. doi:
http://dx.doi.org/10.1016/j.pmatsci.2011.11.001.
[240] Bolzoni L, Esteban PG, Ruiz-Navas EM, Gordo E. Mechanical behaviour of pressed and sintered titanium alloys obtained from master alloy addition
powders. J Mech Behav Biomed Mater 2012;15:33–45. doi: http://dx.doi.org/10.1016/j.jmbbm.2012.05.019.
[241] Li C-L, Ye W-J, Mi X-J, Hui S-X, Lee D-G, Lee Y-T. Development of low cost and low elastic modulus of Ti-Al-Mo-Fe alloys for automotive applications.
In: Imam MA, Froes FHS, Reddy RG, editors. Cost-Affordable Titan Iv, vol. 551. p. 114–7. doi: http://dx.doi.org/10.4028/www.scientific.net/
KEM.551.114.
[242] Bolzoni L, Weissgaerber T, Kieback B, Ruiz-Navas EM, Gordo E. Mechanical behaviour of pressed and sintered CP Ti and Ti-6Al-7Nb alloy obtained
from master alloy addition powder. J Mech Behav Biomed Mater 2013;20:149–61. doi: http://dx.doi.org/10.1016/j.jmbbm.2012.08.022.
[243] Yang F, Zhang D, Gabbitas B, Lu H, Wang C. Microstructural evolution during extrusion of a Ti/Al/Al35V65 (Ti-6Al-4V) powder compact and the
mechanical properties of the extruded rod. Mater Sci Eng a-Struct Mater Prop Microstruct Process 2014;598:360–7. doi: http://dx.doi.org/10.1016/j.
msea.2014.01.055.
[244] Bolzoni L, Herraiz E, Ruiz-Navas EM, Gordo E. Study of the properties of low-cost powder metallurgy titanium alloys by 430 stainless steel addition.
Mater Des 2014;60:628–36. doi: http://dx.doi.org/10.1016/j.matdes.2014.04.019.
[245] StJohn DH, Easton MA, Cao P, Bermingham M, Qian M. A brief history of the grain refinement of cast light alloys. In: Stone I, McKay B, Fan ZY, editors.
Light Met Technol 2013, vol. 765. p. 123–9. doi: http://dx.doi.org/10.4028/www.scientific.net/MSF.765.123.
[246] ASTM International. ASTM F90 - 09 Standard Specification for Wrought Cobalt-20Chromium-15Tungsten-10Nickel Alloy for Surgical Implant
Applications (UNS R30605) 2016. <https://www.astm.org/DATABASE.CART/HISTORICAL/F90-09.htm> [accessed December 15, 2016].
[247] Miyoshi S, Takahashi T, Ohtani M, Yamamoto H, Kameyama K. Analysis of the shape of the tibial tray in total knee arthroplasty using a three
dimension finite element model. Clin Biomech (Bristol, Avon) 2002;17:521–5.
[248] Arsene CTC, Gabrys B. Probabilistic finite element predictions of the human lower limb model in total knee replacement. Med Eng Phys
2013;35:1116–32. doi: http://dx.doi.org/10.1016/j.medengphy.2012.11.011.
[249] Cuilliere J-C, Francois V, Drouet J-M. Towards the integration of topology optimization into the CAD process. Comput Aided Des Appl
2014;11:120–40. doi: http://dx.doi.org/10.1080/16864360.2014.846067.
[250] Tian R, Wu Z, Wang C. Scalable FEA on non-conforming assembly mesh. Comput Meth Appl Mech Eng 2013;266:98–111. doi: http://dx.doi.org/
10.1016/j.cma.2013.07.012.
[251] Kocich R, Kunčická L, Davis CF, Lowe TC, Szurman I, Macháčková A. Deformation behavior of multilayered Al-Cu clad composite during cold-swaging.
Mater Des 2016;90:379–88. doi: http://dx.doi.org/10.1016/j.matdes.2015.10.145.
[252] Lu S, Wang K, Li X, Liu S. A new method for simulating and predicting dynamic recrystallization in metal forging. Acta Metall Sin 2014;50:1128–36.
[253] Kluess D, Souffrant R, Mittelmeier W, Wree A, Schmitz K-P, Bader R. A convenient approach for finite-element-analyses of orthopaedic implants in
bone contact: modeling and experimental validation. Comput Meth Programs Biomed 2009;95:23–30. doi: http://dx.doi.org/10.1016/j.
cmpb.2009.01.004.
[254] La Harrysson O, Hosni YA, Nayfeh JF. Custom-designed orthopedic implants evaluated using finite element analysis of patient-specific computed
tomography data: femoral-component case study. Bmc Musculoskelet Disord 2007;8:0. doi: http://dx.doi.org/10.1186/1471-2474-8-91.
[255] Zhang T, Harrison NM, McDonnell PF, McHugh PE, Leen SB. Micro-macro wear-fatigue of modular hip implant taper-lock coupling. J Strain Anal Eng
Des 2014;49:2–18. doi: http://dx.doi.org/10.1177/0309324713502175.
[256] Liu Y, Zhao X, Wang D. Determination of the plastic properties of materials treated by ultrasonic surface rolling process through instrumented
indentation. Mater Sci Eng a-Structural Mater Prop Microstruct Process 2014;600:21–31. doi: http://dx.doi.org/10.1016/j.msea.2014.01.096.
L. Kunčická et al. / Progress in Materials Science 88 (2017) 232–280 277

[257] Yang C, An Y, Tort M, Hodgson PD. Fabrication, modelling and evaluation of microstructured materials in a digital framework. Comput Mater Sci
2014;81:89–97.
[258] Svyetlichnyy DS. A three-dimensional frontal cellular automaton model for simulation of microstructure evolution—initial microstructure module.
Model Simul Mater Sci Eng 2014;22:85001.
[259] Tang BB, Tang BB, Li J, Zhang F, Yang G. Modeling the high temperature deformation constitutive relationship of TC4-DT alloy based on fuzzy-neural
network. Rare Met Mater Eng 2013;42:1347–51.
[260] Dobatkin SV, Zrnik J, Mamuzic I. Nanostructures by severe plastic deformation of steels: advantages and problems. Metalurgija 2006;45:313–21.
[261] Kocich R, Fiala J, Szurman I, Macháčková A, Mihola M. Twist-channel angular pressing: effect of the strain path on grain refinement and mechanical
properties of copper. J Mater Sci 2011;46:7865–76. doi: http://dx.doi.org/10.1007/s10853-011-5768-1.
[262] Kocich R, Kunická L, Král P, Macháčková A. Sub-structure and mechanical properties of twist channel angular pressed aluminium. Mater Charact
2016;119:75–83. doi: http://dx.doi.org/10.1016/j.matchar.2016.07.020.
[263] Kunčická L, Kocich R, Král P, Pohludka M, Marek M. Effect of strain path on severely deformed aluminium. Mater Lett 2016;180:280–3. doi: http://dx.
doi.org/10.1016/j.matlet.2016.05.163.
[264] Su LH, Lu C, He LZ, Zhang LC, Guagliardo P, Tieu AK, et al. Study of vacancy-type defects by positron annihilation in ultrafine-grained aluminum
severely deformed at room and cryogenic temperatures. Acta Mater 2012;60:4218–28. doi: http://dx.doi.org/10.1016/j.actamat.2012.04.003.
[265] Korznikova E, Schafler E, Steiner G, Zehetbauer MJ. Measurements of vacancy type defects in SPD deformed Ni; 2006.
[266] Lowe TC, Valiev RZ. Frontiers for bulk nanostructured metals in biomedical applications. In: Biomater Biodevices. Wiley Scrivener; 2014. p. 3–52.
[267] Paulin C, Fouvry S, Deyber S. Wear kinetics of Ti-6Al-4V under constant and variable fretting sliding conditions. Wear 2005;259:292–9. doi: http://dx.
doi.org/10.1016/j.wear.2005.01.034.
[268] Long M, Rack H. Titanium alloys in total joint replacement—a materials science perspective. Biomaterials 1998;19:1621–39. doi: http://dx.doi.org/
10.1016/S0142-9612(97)00146-4.
[269] Figueiredo RB, Kawasaki M, Xu C, Langdon TG. Achieving superplastic behavior in fcc and hcp metals processed by equal-channel angular pressing.
Mater Sci Eng a-Struct Mater Prop Microstruct Process 2008;493:104–10. doi: http://dx.doi.org/10.1016/j.msea.2007.06.090.
[270] Liu FC, Ma ZY. Contribution of grain boundary sliding in low-temperature superplasticity of ultrafine-grained aluminum alloys. Scr Mater
2010;62:125–8. doi: http://dx.doi.org/10.1016/j.scriptamat.2009.10.010.
[271] Avtokratova E, Sitdikov O, Markushev M, Mulyukov R. Extraordinary high-strain rate superplasticity of severely deformed Al-Mg-Sc-Zr alloy. Mater
Sci Eng a-Struct Mater Prop Microstruct Process 2012;538:386–90. doi: http://dx.doi.org/10.1016/j.msea.2012.01.041.
[272] Kim WJ, Park IB. Enhanced superplasticity and diffusional creep in ultrafine-grained Mg-6Al-1Zn alloy with high thermal stability. Scr Mater
2013;68:179–82. doi: http://dx.doi.org/10.1016/j.scriptamat.2012.10.011.
[273] Davidson JA, Mishra AK, Kovacs P, Poggie RA. New surface-hardened, low-modulus, corrosion-resistant Ti-13Nb-13Zr alloy for total hip arthroplasty.
Biomed Mater Eng 1994;4:231–43.
[274] Chen S, Jin X, Rong L. Improvement in high temperature oxidation resistance of 9%Cr ferritic–martensitic steel by enhanced diffusion of Mn. Oxid Met
2016;85:189–203. doi: http://dx.doi.org/10.1007/s11085-015-9596-6.
[275] Popescu G, Morales-Espejel GE, Wemekamp B, Gabelli A. An engineering model for three-dimensional elastic-plastic rolling contact analyses. Tribol
Trans 2006;49:387–99.
[276] Ashkenazy Y, Vo NQ, Schwen D, Averback RS, Bellon P. Shear induced chemical mixing in heterogeneous systems. Acta Mater 2012;60:984–93. doi:
http://dx.doi.org/10.1016/j.actamat.2011.11.014.
[277] Arshad SN, Lach TG, Pouryazdan M, Hahn H, Bellon P, Dillon SJ, et al. Dependence of shear-induced mixing on length scale. Scr Mater 2013;68:215–8.
doi: http://dx.doi.org/10.1016/j.scriptamat.2012.10.027.
[278] Kocich R, Greger M, Kursa M, Szurman I, Macháčková A. Twist channel angular pressing (TCAP) as a method for increasing the efficiency of SPD. Mater
Sci Eng A 2010;527:6386–92. doi: http://dx.doi.org/10.1016/j.msea.2010.06.057.
[279] Kunčická L, Lowe TC, Davis CF, Kocich R, Pohludka M. Synthesis of an Al/Al2O3 composite by severe plastic deformation. Mater Sci Eng A
2015;646:234–41. doi: http://dx.doi.org/10.1016/j.msea.2015.08.075.
[280] Rodrigues WC, Broilo LR, Schaeffer L, Knörnschild G, Espinoza FRM. Powder metallurgical processing of Co–28%Cr–6%Mo for dental implants:
physical, mechanical and electrochemical properties. Powder Technol 2011;206:233–8. doi: http://dx.doi.org/10.1016/j.powtec.2010.09.024.
[281] Haase C, Ng HP, Lapovok R, Estrin Y. Ti-6Al-4V billet produced by compaction of be powders using equal-channel angular pressing. In: Qian M, editor.
Powder metall titan powder process consol metall titan, vol. 520. p. 301–8. doi: http://dx.doi.org/10.4028/www.scientific.net/KEM.520.301.
[282] Luo SD, Yang YF, Schaffer GB, Qian M. Warm die compaction and sintering of titanium and titanium alloy powders. J Mater Process Technol
2014;214:660–6. doi: http://dx.doi.org/10.1016/j.jmatprotec.2013.10.010.
[283] Yang YF, Imai H, Kondoh K, Qian M. Comparison of spark plasma sintering of elemental and master alloy powder mixes and prealloyed Ti-6Al-4V
powder. Int J Powder Metall 2014;50:41–7.
[284] Guden M, Celik E, Akar E, Cetiner S. Compression testing of a sintered Ti6Al4V powder compact for biomedical applications. Mater Charact
2005;54:399–408. doi: http://dx.doi.org/10.1016/j.matchar.2005.01.006.
[285] Bardos D. High strength Co-Cr-Mo alloy by hot isostatic pressing of powder. Biomater Med Devices Artif Organs 1979;7:73–80. doi: http://dx.doi.org/
10.3109/10731197909119373.
[286] El-Soudani SM, Yu K-O, Crist EM, Sun F, Campbell MB, Esposito TS, et al. Optimization of blended-elemental powder-based titanium alloy extrusions
for aerospace applications. Metall Mater Trans a-Phys Metall Mater Sci 2013;44A:899–910. doi: http://dx.doi.org/10.1007/s11661-012-1437-5.
[287] Bolzoni L, Babu NH, Ruiz-Navas EM, Gordo E. Comparison of microstructure and properties of Ti-6Al-7Nb alloy processed by different powder
metallurgy routes. In: Imam MA, Froes FHS, Reddy RG, editors. Cost-Affordable Titan Iv, vol. 551. p. 161–79. doi: http://dx.doi.org/10.4028/
www.scientific.net/KEM.551.161.
[288] Kurgan N, Varol R. Mechanical properties of P/M 316L stainless steel materials. Powder Technol 2010;201:242–7. doi: http://dx.doi.org/10.1016/j.
powtec.2010.03.041.
[289] Chen Y, Li Y, Kurosu S, Yamanaka K, Tang N, Chiba A. Effects of microstructures on the sliding behavior of hot-pressed CoCrMo alloys. Wear
2014;319:200–10. doi: http://dx.doi.org/10.1016/j.wear.2014.07.022.
[290] Manukyan K, Amirkhanyan N, Aydinyan S, Danghyan V, Grigoryan R, Sarkisyan N, et al. Novel NiZr-based porous biomaterials: synthesis and in vitro
testing. Chem Eng J 2010;162:406–14. doi: http://dx.doi.org/10.1016/j.cej.2010.05.042.
[291] Kurgan N. Effect of porosity and density on the mechanical and microstructural properties of sintered 316L stainless steel implant materials. Mater
Des 2014;55:235–41. doi: http://dx.doi.org/10.1016/j.matdes.2013.09.058.
[292] Yang X, Hutchinson CR. Corrosion-wear of b-Ti alloy TMZF (Ti-12Mo-6Zr-2Fe) in simulated body fluid. Acta Biomater 2016;42:429–39. doi: http://dx.
doi.org/10.1016/j.actbio.2016.07.008.
[293] Sidambe AT. Biocompatibility of advanced manufactured titanium implants-a review. Materials (Basel) 2014;7:8168–88. doi: http://dx.doi.org/
10.3390/ma7128168.
[294] Vaezi M, Seitz H, Yang S. A review on 3D micro-additive manufacturing technologies. Int J Adv Manuf Technol 2012;67:1721–54. doi: http://dx.doi.
org/10.1007/s00170-012-4605-2.
[295] Maleksaeedi S, Wang JK, El-Hajje A, Harb L, Guneta V, He Z, et al. Toward 3D printed bioactive titanium scaffolds with bimodal pore size distribution
for bone ingrowth. Proc CIRP 2013;5:158–63. doi: http://dx.doi.org/10.1016/j.procir.2013.01.032.
[296] Li JP, de Wijn JR, Van Blitterswijk Ca, de Groot K. Porous Ti6Al4V scaffold directly fabricating by rapid prototyping: preparation and in vitro
experiment. Biomaterials 2006;27:1223–35. doi: http://dx.doi.org/10.1016/j.biomaterials.2005.08.033.
278 L. Kunčická et al. / Progress in Materials Science 88 (2017) 232–280

[297] Karlsson J, Sjogren T, Snis A, Engqvist H, Lausmaa J. Digital image correlation analysis of local strain fields on Ti6Al4V manufactured by electron beam
melting. Mater Sci Eng a-Struct Mater Prop Microstruct Process 2014;618:456–61. doi: http://dx.doi.org/10.1016/j.msea.2014.09.022.
[298] Shi T, Lu B, Shi S, Meng W, Fu G. Laser metal deposition with spatial variable orientation based on hollow-laser beam with internal powder feeding
technology. Opt Laser Technol 2017;88:234–41. doi: http://dx.doi.org/10.1016/j.optlastec.2016.09.019.
[299] Yadroitsev I, Krakhmalev P, Yadroitsava I. Selective laser melting of Ti6Al4V alloy for biomedical applications: temperature monitoring and
microstructural evolution. J Alloys Compd 2014;583:404–9. doi: http://dx.doi.org/10.1016/j.jallcom.2013.08.183.
[300] Tang Q, Pang S, Chen B, Suo H, Zhou J. A three dimensional transient model for heat transfer and fluid flow of weld pool during electron beam freeform
fabrication of Ti-6-Al-4-V alloy. Int J Heat Mass Transf 2014;78:203–15. doi: http://dx.doi.org/10.1016/j.ijheatmasstransfer.2014.06.048.
[301] Gaytan SM, Murr LE, Martinez E, Martinez JL, Machado BI, Ramirez DA, et al. Comparison of microstructures and mechanical properties for solid and
mesh cobalt-base alloy prototypes fabricated by electron beam melting. Metall Mater Trans A 2010;41:3216–27. doi: http://dx.doi.org/10.1007/
s11661-010-0388-y.
[302] Viswanathan V, Laha T, Balani K, Agarwal A, Seal S. Challenges and advances in nanocomposite processing techniques. Mater Sci Eng R Reports
2006;54:121–285. doi: http://dx.doi.org/10.1016/j.mser.2006.11.002.
[303] El-Hajje A, Kolos EC, Wang JK, Maleksaeedi S, He Z, Wiria FE, et al. Physical and mechanical characterisation of 3D-printed porous titanium for
biomedical applications. J Mater Sci Mater Med 2014;25:2471–80. doi: http://dx.doi.org/10.1007/s10856-014-5277-2.
[304] Bandyopadhyay A, Espana F, Balla VK, Bose S, Ohgami Y, Davies NM. Influence of porosity on mechanical properties and in vivo response of Ti6Al4V
implants. Acta Biomater 2010;6:1640–8. doi: http://dx.doi.org/10.1016/j.actbio.2009.11.011.
[305] Verlee B, Dormal T, Lecomte-Beckers J. Density and porosity control of sintered 316L stainless steel parts produced by additive manufacturing.
Powder Metall 2012;55:260–7. doi: http://dx.doi.org/10.1179/0032589912z.00000000082.
[306] Carreno-Morelli E, Martinerie S, Mucks L, Cardis B. Three-dimensional printing of stainless steel parts. In: Salgado L, filho FA, editors. Adv Powder
Technol Vi, vols. 591–593. p. 374–9.
[307] Hunt JA, Callaghan JT, Sutcliffe CJ, Morgan RH, Halford B, Black RA. The design and production of Co-Cr alloy implants with controlled surface
topography by CAD-CAM method and their effects on osseointegration. Biomaterials 2005;26:5890–7. doi: http://dx.doi.org/10.1016/j.
biomaterials.2005.03.004.
[308] Reclaru L, Ardelean L, Rusu L, Sinescu C. Co-Cr material selection in prosthetic restoration: laser sintering technology. In: Nicoara M, Raduta A, Opris C,
editors. Adv Mater Struct Iv, vol. 188. p. 412–5. doi: http://dx.doi.org/10.4028/www.scientific.net/SSP.188.412.
[309] Samuel S, Nag S, Scharf TW, Banerjee R. Wear resistance of laser-deposited boride reinforced Ti-Nb–Zr–Ta alloy composites for orthopedic implants.
Mater Sci Eng C 2008;28:414–20. doi: http://dx.doi.org/10.1016/j.msec.2007.04.029.
[310] Banerjee R, Nag S, Samuel S, Fraser HL. Laser-deposited Ti-Nb-Zr-Ta orthopedic alloys. J Biomed Mater Res Part A 2006;78A:298–305. doi: http://dx.
doi.org/10.1002/jbm.a.30694.
[311] Gräfe W. Quantum mechanical models of metal surfaces and nanoparticles. Springer; 2015.
[312] DrugWatch. Hip Replacement Recall – Implant Failures & FDA Warnings 2012. <https://www.drugwatch.com/hip-replacement/recalls/> [accessed
December 16, 2016].
[313] Chu PK. Surface engineering and modification of biomaterials. Thin Solid Films 2013;528:93–105. doi: http://dx.doi.org/10.1016/j.tsf.2012.07.144.
[314] Vrancken B, Thijs L, Kruth J-P, Van Humbeeck J. Heat treatment of Ti6Al4V produced by selective laser melting: microstructure and mechanical
properties. J Alloys Compd 2012;541:177–85. doi: http://dx.doi.org/10.1016/j.jallcom.2012.07.022.
[315] Zhang F, Reich M, Kessler O, Burkel E. The potential of rapid cooling spark plasma sintering for metallic materials. Mater Today 2013;16:192–7. doi:
http://dx.doi.org/10.1016/j.mattod.2013.05.005.
[316] Chinese Doctors Successfully Use 3D Printing for Major Hip Replacement Surgery | 3DPrint.com n.d.
[317] 3D-Printed Custom-Made Hip Joint Held in Place with Stem Cells | Software & Services for Biomedical Engineering n.d.
[318] Bagherifard S, Hickey DJ, de Luca AC, Malheiro VN, Markaki AE, Guagliano M, et al. The influence of nanostructured features on bacterial adhesion and
bone cell functions on severely shot peened 316L stainless steel. Biomaterials 2015;73:185–97. doi: http://dx.doi.org/10.1016/j.
biomaterials.2015.09.019.
[319] Zhang H, Hei Z, Liu G, Lu J, Lu K. Formation of nanostructured surface layer on AISI 304 stainless steel by means of surface mechanical attrition
treatment. Acta Mater 2003;51:1871–81. doi: http://dx.doi.org/10.1016/S1359-6454(02)00594-3.
[320] Bagherifard S, Guagliano M. Review of shot peening processes to obtain nanocrystalline surfaces in metal alloys. Surf Eng 2009;25:3–14. doi: http://
dx.doi.org/10.1179/026708408x334087.
[321] Shekhar S, Abolghasem S, Basu S, Cai J, Shankar MR. Effect of severe plastic deformation in machining elucidated via rate-strain-microstructure
mappings. J Manuf Sci Eng Asme 2012;134. doi: http://dx.doi.org/10.1115/1.4006549.
[322] Iglesias P, Moscoso W, Mann JB, Saldana C, Shankar MR, Chandrasekar S, et al. Production analysis of new machining-based deformation processes for
nanostructured materials. Int J Mater Form 2008;1:459–62. doi: http://dx.doi.org/10.1007/s12289-008-0094-0.
[323] Saldana C, Yang P, Mann JB, Moscoso W, Gill DD, Chandrasekar S, et al. Micro-scale components from high-strength nanostructured alloys. Mater Sci
Eng a-Struct Mater Prop Microstruct Process 2009;503:172–5. doi: http://dx.doi.org/10.1016/j.msea.2008.02.056.
[324] Mann JB, Guo Y, Saldana C, Yeung H, Compton WD, Chandrasekar S. Modulation-assisted machining: a new paradigm in material removal processes.
Adv Mater Res 2011;223:514–22. doi: http://dx.doi.org/10.4028/www.scientific.net/AMR.223.514.
[325] Mann JB, Saldana C, Chandrasekar S, Compton WD, Trumble KP. Metal particulate production by modulation-assisted machining. Scr Mater
2007;57:909–12. doi: http://dx.doi.org/10.1016/j.scriptamat.2007.07.025.
[326] Jagielski J, Piatkowska A, Aubert P, Thomé L, Turos A, Kader AA. Ion implantation for surface modification of biomaterials. Surf Coat Technol
2005;200:6355–61. doi: http://dx.doi.org/10.1016/j.surfcoat.2005.11.005.
[327] Tsyganov I, Lode A, Hanke T, Kolitsch A, Gelinsky M. Osteoblast responses to novel titanium-based surfaces produced by plasma-and ion beam
technologies; 2013. doi:http://dx.doi.org/10.1039/c3ra23351k.
[328] Yamada I, Matsuo J, Toyoda N, Aoki T, Seki T. Progress and applications of cluster ion beam technology. Curr Opin Solid State Mater Sci 2015. doi:
http://dx.doi.org/10.1016/j.cossms.2014.11.002.
[329] García JA, Rodríguez RJ. Ion implantation techniques for non-electronic applications. Vaccum 2011. doi: http://dx.doi.org/10.1016/
j.vacuum.2010.12.024.
[330] Conrad JR, Radtke JL, Dodd RA, Worzala FJ, Tran NC, Tran N. Plasma source ion-implantation technique for surface modification of materials. J Appl
Phys 1987. doi: http://dx.doi.org/10.1063/1.1653549.
[331] García JA, Díaz C, Mändl S, Lutz J, Martínez R, Rodríguez RJ. Tribological improvements of plasma immersion implanted CoCr alloys. Surf Coat Technol
2010;204:2928–32. doi: http://dx.doi.org/10.1016/j.surfcoat.2010.03.050.
[332] Maendl S, Diaz C, Gerlach JW, Garcia JA. Near surface analysis of duplex PIII treated CoCr alloys. Nucl Instrum Meth Phys Res Sect B-Beam Interact
with Mater Atoms 2013;307:305–9. doi: http://dx.doi.org/10.1016/j.nimb.2012.11.052.
[333] Maendl S, Lutz J, Diaz C, Gerlach JWW, Garcia JA, Mändl S, et al. Influence of reduced current density on diffusion and phase formation during PIII
nitriding of austenitic stainless steel and CoCr alloys. Surf Coat Technol 2014;239:116–22. doi: http://dx.doi.org/10.1016/j.surfcoat.2013.11.029.
[334] Galdikas A, Petraitiene A, Moskalioviene T. Internal stress assisted nitrogen diffusion in plasma nitrided medical CoCr alloys. Vacuum
2015;119:233–8. doi: http://dx.doi.org/10.1016/j.vacuum.2015.05.029.
[335] Alonso F, Ugarte JJ, Sansom D, Viviente JL, Oñate JI. Effects of ion implantation on Ti-6Al-4V on its frictional behaviour against UHMWPE. Surf Coat
Technol 1996;83:301–6. doi: http://dx.doi.org/10.1016/0257-8972(96)02767-3.
[336] Paital SR, Dahotre NB. Calcium phosphate coatings for bio-implant applications: materials, performance factors, and methodologies. Mater Sci Eng R
Reports 2009;66:1–70. doi: http://dx.doi.org/10.1016/j.mser.2009.05.001.
L. Kunčická et al. / Progress in Materials Science 88 (2017) 232–280 279

[337] Feng K, Wang Y, Li Z, Chu PK. Characterization of carbon ion implantation induced graded microstructure and phase transformation in stainless steel.
Mater Charact 2015;106:11–9. doi: http://dx.doi.org/10.1016/j.matchar.2015.04.019.
[338] Campbell F. Elements of metallurgy and engineering alloys; 2008. doi:http://dx.doi.org/10.1361/emea2008p243.
[339] Christiansen TL, Somers MAJ. Low-temperature gaseous surface hardening of stainless steel: the current status. Int J Mater Res 2009;100:1361–77.
doi: http://dx.doi.org/10.3139/146.110202.
[340] Luo HS, Zhao C. Low temperature salt bath hardening of AISI 201 austenitic stainless steel. Phys Proc 2013;50:38–42. doi: http://dx.doi.org/10.1016/j.
phpro.2013.11.008.
[341] Surface hardening of steels: understanding the basics. ASM International; 2002.
[342] Wang Q, Huang C, Zhang L. Microstructure and tribological properties of plasma nitriding cast CoCrMo alloy. J Mater Sci Technol 2012;28:60–6. doi:
http://dx.doi.org/10.1016/S1005-0302(12)60024-3.
[343] Manova D, Höche T, Mändl S, Neumann H. Development of CrN precipitates during the initial stages of PIII nitriding of stainless steel thin films. Nucl
Instrum Meth Phys Res Sect B Beam Interact with Mater Atoms 2009;267:1536–9. doi: http://dx.doi.org/10.1016/j.nimb.2009.01.075.
[344] Lutz J, Díaz C, García JA, Blawert C, Mändl S. Corrosion behaviour of medical CoCr alloy after nitrogen plasma immersion ion implantation. Surf Coat
Technol 2011;205:3043–9. doi: http://dx.doi.org/10.1016/j.surfcoat.2010.11.017.
[345] Gutmanas EY, Gotman I. Protective coatings on medical implants by reactive diffusion. Int Conf Phys Mesomech MULTILEVEL Syst 2014, vol. 1623. AIP
Publishing; 2014. p. 203–8. doi: http://dx.doi.org/10.1063/1.4901481.
[346] Samsonov GV. Diffusion cladding of metals, vol. 3. Springer Science & Business Media; 2012.
[347] Park J-W, Kim Y-J, Park CH, Lee D-H, Ko YG, Jang J-H, et al. Enhanced osteoblast response to an equal channel angular pressing-processed pure
titanium substrate with microrough surface topography. Acta Biomater 2009;5:3272–80. doi: http://dx.doi.org/10.1016/j.actbio.2009.04.038.
[348] Luong-Van E, Rodriguez I, Low HY, Elmouelhi N, Lowenhaupt B, Natarajan S, et al. Review: micro- and nanostructured surface engineering for
biomedical applications. J Mater Res 2012;28:165–74. doi: http://dx.doi.org/10.1557/jmr.2012.398.
[349] Scopelliti PE, Borgonovo A, Indrieri M, Giorgetti L, Bongiorno G, Carbone R, et al. The effect of surface nanometre-scale morphology on protein
adsorption. PLoS ONE 2010;5:e11862. doi: http://dx.doi.org/10.1371/journal.pone.0011862.
[350] Elter P, Lange R, Beck U. Atomic force microscopy studies of the influence of convex and concave nanostructures on the adsorption of fibronectin.
Colloids Surf B Biointerfaces 2012;89:139–46. doi: http://dx.doi.org/10.1016/j.colsurfb.2011.09.021.
[351] Kortge A, Elter P, Lange R, Beck U. Simulation of the electric field distribution near a topographically nanostructured titanium-electrolyte interface:
influence of the passivation layer. J Nanomater 2013. doi: http://dx.doi.org/10.1155/2013/820914.
[352] Wang G, Liu X, Zreiqat H, Ding C. Enhanced effects of nano-scale topography on the bioactivity and osteoblast behaviors of micron rough ZrO(2)
coatings. Colloids Surf B-Biointerfaces 2011;86:267–74. doi: http://dx.doi.org/10.1016/j.colsurfb.2011.04.006.
[353] Dallmiglio M, Schaaff P, Holzwarth U, Chiesa R, Rondelli G. The effect of surface treatments on the fretting behavior of Ti-6Al-4V alloy. J Biomed Mater
Res Part B-Appl Biomater 2008;86B:407–16. doi: http://dx.doi.org/10.1002/jbm.b.31034.
[354] Zhang E, Wang Y, Gao F, WeiD S, Zheng Y. Enhanced bioactivity of sandblasted and acid-etched titanium surfaces. In: Yin YSWX, editor. Multi-
Functional Mater Struct Ii, Pts 1 2, vols. 79–82. p. 393–6. doi: http://dx.doi.org/10.4028/www.scientific.net/AMR.79-82.393.
[355] Li BE, Li Y, Li J, Fu XL, Li HP, Wang HS, et al. Influence of nanostructures on the biological properties of Ti implants after anodic oxidation. J Mater Sci
Med 2014;25:199–205. doi: http://dx.doi.org/10.1007/s10856-013-5064-5.
[356] Nieto A, Zhao JM, Han Y-H, Hwang KH, Schoenung JM. Microscale tribological behavior and in vitro biocompatibility of graphene nanoplatelet
reinforced alumina. J Mech Behav Biomed Mater 2016;61:122–34. doi: http://dx.doi.org/10.1016/j.jmbbm.2016.01.020.
[357] Dorner-Reisel A, Schurer C, Klemm V, Irmer G, Muller E. Nano- and microstructure of diamond-like carbon films modified by Ca-O incorporation.
Diam Relat Mater 2003;12:1030–3. doi: http://dx.doi.org/10.1016/s0925-9635(02)00293-5.
[358] Allen M, Myer B, Rushton N. In vitro and in vivo investigations into the biocompatibility of diamond-like carbon (DLC) coatings for orthopedic
applications. J Biomed Mater 2001.
[359] Cui F, Luo Z. Biomaterials modification by ion-beam processing. Surf Coat Technol 1999;112:278–85. doi: http://dx.doi.org/10.1016/S0257-8972(98)
00763-4.
[360] Vacik J, Lavrentiev V, Novotna K, Bacakova L, Lisa V, Vorlicek V, et al. Fullerene (C-60)-transitional metal (Ti) composites: structural and biological
properties of the thin films. Diam Relat Mater 2010;19:242–6. doi: http://dx.doi.org/10.1016/j.diamond.2009.10.016.
[361] Abdi S, Oswald S, Gostin PF, Helth A, Sort J, Baró MD, et al. Designing new biocompatible glass-forming Ti 75- x Zr 10 Nb x Si 15 (x = 0, 15) alloys:
corrosion, passivity, and apatite formation. J Biomed Mater Res Part B Appl Biomater 2016;104:27–38. doi: http://dx.doi.org/10.1002/jbm.b.33332.
[362] Lewis G. Properties of open-cell porous metals and alloys for orthopaedic applications. J Mater Sci Med 2013;24:2293–325. doi: http://dx.doi.org/
10.1007/s10856-013-4998-y.
[363] Baoe L, Xuanyong L, Fanhao M, Jiang C, Chuanxian D. Preparation and antibacterial properties of plasma sprayed nano-titania/silver coatings. Mater
Chem Phys 2009:118. doi: http://dx.doi.org/10.1016/j.matchemphys.2009.07.011.
[364] Zhao C, Cao P, Ji W, Han P, Zhang J, Zhang F, et al. Hierarchical titanium surface textures affect osteoblastic functions. J Biomed Mater Res Part A
2011;99A:666–75. doi: http://dx.doi.org/10.1002/jbm.a.33239.
[365] Jiang JY, Xu JL, Liu ZH, Deng L, Sun B, Liu SD, et al. Preparation, corrosion resistance and hemocompatibility of the superhydrophobic TiO2 coatings on
biomedical Ti-6Al-4V alloys. Appl Surf Sci 2015;347:591–5. doi: http://dx.doi.org/10.1016/j.apsusc.2015.04.075.
[366] Feruz Y, Mordehai D. Towards a universal size-dependent strength of face-centered cubic nanoparticles. Acta Mater 2016;103:433–41. doi: http://dx.
doi.org/10.1016/j.actamat.2015.10.027.
[367] Zhu T, Li J. Ultra-strength materials. Prog Mater Sci 2010;55:710–57. doi: http://dx.doi.org/10.1016/j.pmatsci.2010.04.001.
[368] Isaacs RA, Zhu H, Preston C, Mansour A, LeMieux M, Zavalij PY, et al. Nanocarbon-copper thin film as transparent electrode. Appl Phys Lett
2015;106:193108. doi: http://dx.doi.org/10.1063/1.4921263.
[369] Salamanca-Riba LG, Isaacs RA, LeMieux MC, Wan J, Gaskell K, Jiang Y, et al. Synthetic crystals of silver with carbon: 3D epitaxy of carbon
nanostructures in the silver lattice. Adv Funct Mater 2015;25:4768–77. doi: http://dx.doi.org/10.1002/adfm.201501156.
[370] Naeimirad M, Zadhoush A, Neisiany RE. Fabrication and characterization of silicon carbide/epoxy nanocomposite using silicon carbide nanowhisker
and nanoparticle reinforcements. J Compos Mater 2015;50:435–46. doi: http://dx.doi.org/10.1177/0021998315576378.
[371] Langer M, Peyrin F. 3D X-ray ultra-microscopy of bone tissue. Osteoporos Int 2016;27:441–55. doi: http://dx.doi.org/10.1007/s00198-015-3257-0.
[372] Talamonti C, Reggioli V, Bruzzi M, Bucciolini M, Civinini C, Marrazzo L, et al. Proton radiography for clinical applications. Nucl Instrum Meth Phys Res
Sect A Accel Spectrometers, Detect Assoc Equip 2010;612:571–5. doi: http://dx.doi.org/10.1016/j.nima.2009.08.040.
[373] Dalecki D, Hocking DC. Ultrasound technologies for biomaterials fabrication and imaging. Ann Biomed Eng 2014. doi: http://dx.doi.org/10.1007/
s10439-014-1158-6.
[374] Migliori A, Sarrao J, Visscher W. Resonant ultrasound spectroscopic techniques for measurement of the elastic moduli of solids. Phys B Condens
Matter 1993;183:1–24.
[375] Schwarz R, Vuorinen J. Resonant ultrasound spectroscopy: applications, current status and limitations. J Alloys Compd 2000;310:243–50. doi: http://
dx.doi.org/10.1016/S0925-8388(00)00925-7.
[376] Maynard J. Resonant ultrasound spectroscopy. Med Imaging’98; 1998.
[377] Haupert S, Guérard S, Peyrin F, Mitton D, Laugier P. Non destructive characterization of cortical bone micro-damage by nonlinear resonant ultrasound
spectroscopy. PLoS ONE 2014.
[378] Bernard S, Schneider J, Varga P. Elasticity–density and viscoelasticity–density relationships at the tibia mid-diaphysis assessed from resonant
ultrasound spectroscopy measurements. Biomech Model Mechanobiol 2015.
280 L. Kunčická et al. / Progress in Materials Science 88 (2017) 232–280

[379] Kinney J, Gladden J, Marshall G. Resonant ultrasound spectroscopy measurements of the elastic constants of human dentin. J Biomech 2004.
[380] Ganesan V, Laha K, Mathew MD. Influence of nitrogen content on the evolution of creep damage in 316 LN stainless steel. Proc Eng 2014;86:58–65.
doi: http://dx.doi.org/10.1016/j.proeng.2014.11.011.
[381] Hollander DA, von Walter M, Wirtz T, Sellei R, Schmidt-Rohlfing B, Paar O, et al. Structural, mechanical and in vitro characterization of individually
structured Ti-6Al-4V produced by direct laser forming. Biomaterials 2006;27:955–63. doi: http://dx.doi.org/10.1016/j.biomaterials.2005.07.041.

You might also like