You are on page 1of 10

Colloids and Surfaces B: Biointerfaces 39 (2004) 133142

Biomaterials in total joint replacement


Kalpana S. Katti
Department of Civil Engineering, North Dakota State University, CIE 201B, Fargo, ND 58105, USA Available online 20 February 2004

Abstract The current state of materials systems used in total hip replacement is presented in this paper. An overview of the various material systems used in total hip replacement reported in literature is presented in this paper. Metals, polymers, ceramics and composites are used in the design of the different components of hip replacement implants. The merits and demerits of these material systems are evaluated in the context of mechanical properties most suitable for total joint replacement such as a hip implant. Current research on advanced polymeric nanocomposites and biomimetic composites as novel materials systems for bone replacement is also discussed. This paper examines the current research in the materials science and the critical issues and challenges in these materials systems that require further research before application in biomedical industry. 2003 Elsevier B.V. All rights reserved.
Keywords: THR; Biomedical; Review; Bone; Mechanical properties

1. Introduction 1.1. Natural bone structure and mechanical properties Natural bone is a composite material made up of collagen ber matrix stiffened by hydroxyapatite (HAP) (Ca10 (PO4 )6 (OH)2 ) crystals that account for 69% of the weight of the bone [1]. The organic phase is composed mainly of a protein, type I collagen. Like all collagens the type I collagen is a triple helix. The protein molecule, tropocollagen in the organic phase of bone is 260 nm long and the molecules alongside each other are staggered by about 1/4 of their length [2]. Histologically, the bone is divided into immature woven bone and mature lamellar bone. In woven bone, the collagen is ne bered, 0.1 m in diameter, and is oriented almost randomly. The woven bone consists of cells (osteocytes) and blood vessels. The collagen in lamellar bone forms branching bundles, 23 m in diameter. Collagen-based biomaterials are routinely used as sutures, blood vessels, heart valves and drug delivery systems [3,4]. Generally, elastic modulus, tensile and compressive strengths are the mechanical properties investigated to ensure suitability of the biomaterial. Mechanical properties of bone are shown in Table 1. The elastic modulus of

bone (17 GPa in tension in human femur) is intermediate between that of apatite and collagen [5]. But its strength is higher than that of both. The organic phase behaves as a compliant material with high toughness. The inorganic phase, HAP, is present in the form of small crystallites of dimensions 5 nm 20 nm 40 nm. The stiffness of this material is about two-thirds of steel. Also, it is quite brittle and has poor impact resistance and fractures easily. The properties of bone really arise from combination of the high hardness (of HAP) and high fracture toughness (of organic). This superposition of two very dissimilar materials with entirely different properties results in the formation of a nanocomposite system of which the physical properties surpass that of the individual components. Thus, bone represents a bio-nanocomposite system that has evolved over millions of years and perfected with optimized properties. Bone can remodel and adapt itself to the applied mechanical environment. This property of bone that is also called as Wolffs law, results such that the new remodeled structure is more suitably adapted to the applied load. Further the application of higher stress results in a more dense bone. 1.2. Materials consideration for implants Biomaterials were rst dened as nonviable materials used in a medical device, intended to interact with biological systems [6]. Further, Black dened the term biomateri-

Tel.: +1-701-231-9504; fax: +1-701-231-6185. E-mail address: Kalpana.katti@ndsu.nodak.edu (K.S. Katti).

0927-7765/$ see front matter 2003 Elsevier B.V. All rights reserved. doi:10.1016/j.colsurfb.2003.12.002

134

K.S. Katti / Colloids and Surfaces B: Biointerfaces 39 (2004) 133142

Nomenclature BCP SEVA HAP HDPE UHMWPE PA PS PE PP PU PPh PEEK PTFE PET PMMA PGA PTC DLPLA PDO DLPLLA DLPLG PHA PGA-TMC SR LPLG PCL THR TCP biphasic calcium phosphate ethylene vinyl alcohol copolymer hydroxyapatite high density polyethylene ultrahigh molecular weight polyethylene polyacetal polysulfone polyethylene polypropylene polyurethane polyphosphosone polyetheretherketone polytetrauoroethylene polyethylene terepthalate polymethylmethacrylate poly(glycolide) poly(trimethylene carbonate) poly(dl-lactide) poly(dioxanone) poly(dl-lactide-co-l-lactide) poly(dl-lactide-co-glycolide) poly(-hydroxyalkanoates) poly(glycolide-co-trimethylene carbonate) silicone rubber poly(l-lactide-co-glycolide) poly(-caprolactone) total hip replacement tricalcium phosphate

compatibility is only a qualitative description of how the body tissues interact with the biomaterial within some expectations of certain implantation purpose and site [8]. The average load on a hip joint is estimated to be up to three times body weight and the peak load during other strenuous activities such as jumping can be as high as 10 times body weight. In addition hip bones are subjected to cyclic loading as high as 106 cycles in 1 year [9]. Materials scientists have investigated metals, ceramics, polymers and composites as biomaterials. The general criteria for materials selection for bone implant materials are: It is highly biocompatible and does not cause an inammatory or toxic response beyond an acceptable tolerable level. It has appropriate mechanical properties, closest to bone. Manufacturing and processing methods are economically viable. Ideally, a bone implant such as a hip implant should be such that it exhibits an identical response to loading as real bone and is also biocompatible with existing tissue. The compatibility issue involves surface compatibility, mechanical compatibility and also osteocompatibility. These materials are also classied as bioactive (illicit a favorable response from tissue and bond well), bioinert and biodegradable. In this paper, the different materials and composites used in the fabrication of implants such as total hip replacement are investigated for their merits and demerits. The purpose of this paper is to provide a review of the various material systems currently being investigated as potential components of the total hip replacement (THR) implants. Replacement of joints such as a THR is a serious health concern. In the USA alone, over 150,000 total hip replacement procedures are undertaken annually. Most hip implants must be replaced after 15 years. It is estimated that 20% of hip replacement surgeries simply replace the original, failed implant. Of all the joints in a human body, the hip and knee represents some of the most important synovial joints. The hip joint consists of two complementary articular surfaces separated by articular cartilage and the synovial uid that has a pH between 7.29 and 7.45. Excessive wear of the interfaces due to degenerative disease (such as osteoarthritis) or injury requires a replacement of the entire hip joint. Historically, a total hip replacement the articulation of a human hip is simulated with the use of two components, a cup type and a long femoral type element. A typical hip implant fabricated from titanium is shown in Fig. 1. The head of the femoral element ts inside the cup to enable the articulation of human joint. These two parts of the hip implant have been made using a variety of materials such as metals, ceramics, polymers and composites. Typically polymeric materials alone tend to be too weak to be suitable for meeting the requirement of stress deformation responses in the THR components. Metals typically have good mechanical properties but show poor biocompatibility, cause stress shielding and release of dangerous metal ions causing eventual failure and removal of

als as materials of natural or manmade origin that are used to direct, supplement, or replace the functions of living tissues [7]. Many synthetic materials are used in the medicine for a variety of applications ranging from total replacement of hard or soft tissues (such as bone plates, pins, total joint replacement, dental implants, intra-ocular lenses, etc.), repair, diagnostic or corrective devices (such as pacemakers, catheters, heart valves, etc.). The two primary issues in materials science of new bone biomaterials are mechanical properties and biocompatibility. Although mechanical properties of biomaterials have been well characterized, the term bioTable 1 Mechanical properties of bones (adapted from [2,5,7]) Hard tissues Tibia Femur Radius Humerus Cervical Lumbar Compressive strength (MPa) 159 167 114 132 10 5 Tensile strength (MPa) 140 121 149 130 3.1 3.7 Elastic modulus (GPa) 18.1 17.2 18.6 17.2 0.23 0.16

K.S. Katti / Colloids and Surfaces B: Biointerfaces 39 (2004) 133142

135

ical interlock, subsequent bone resorption often results due to the modulus mismatch between cancellous bone and cement. Fixation of the implants with polymethymethacrylate (PMMA) allows patients to bear weight instantly as opposed to a wait of about 12 weeks for implants that are attached by only mechanical interlocking with bone ingrowth. Typically implants are roughened and coated with PMMA before applying bone cement (also PMMA). The bone cement interface is highly dynamic with degradation of the polymer in the cement and bone ingrowth. The nature of this interface is specic to the materials used in implants. The following sections evaluate the different materials systems used in orthopedic applications, particularly for total hip replacement.

2. Material systems in total hip replacement 2.1. Metals


Fig. 1. A typical hip implant with titanium femoral (a) and polyethylene acetabular cup (b) components.

implant. Ceramics generally have good biocompatibility but poor fracture toughness and tend to be brittle. Composite materials with engineered interfaces resulting in combination of biocompatibility, mechanical strength and toughness, is the focus of many current studies. Total joint replacements generally involve implantation components held in place by a cement. Loosening of the components often occurs at the interface between the cement and bone due to failure of the xation of the cement to the bone. Although deeper penetration of the cement into the interstices of cancellous bone should improve the mechan-

The effort to nd substitutions for repair of seriously damaged human bones dates back to centuries. Metals have been the primary materials in the past for this purpose due to their superior mechanical properties [10], albeit dangerous ions that are released in vivo from these alloys. Originally femoral components of the THR were made of stainless steel that was replaced by a cobalt-chromium-molybdenum alloy (VitalliumTM ) [11,12]. Mechanical properties of common metallic THR materials are shown in Table 2. Metallurgical heat treatments and resulting microstructures guide the resulting mechanical properties in metallic implant materials [14]. Most commonly, the long femoral element is made of stainless steel, CoCr alloys, or Ti alloys, and the cup component is made up of alumina or zirconia ceramic, polytertrauoro ethylene (PTFE) or CoCr alloy.

Table 2 Mechanical properties of alloys in total joint replacement (adapted from [10,13]) Alloy cpTi (pure titanium) TiZr CoCr alloys CoCrMo Ti6Al4V Ti6Al7Nb Ti5Al2.5Fe Ti13Nb13Zr Ti15Mo5Zr3Al Ti12Mo6Zr2Fe Ti15Mo5Zr3Al Stainless steel 316 L Ti15Mo2.8Nb3Al Ti35Nb5Ta7Zr (TNZT) Ti15Mo3Nb0.3O (21SRx) Ti35Nb5Ta7Zr0.4O (TNZTO) Ti0/20Zr0/20Sn4/8Nb2/4Ta+(Pd, N, O) Microstructure { } Cast { /} {Austenite (fcc) + hcp} { /} {/} {/} { /} {Metastable } {Aged + } {Metastable } {Metastable } {Aged + } {Austenite} {Metastable } {Aged + } {Metastable } {Metastable } + silicides {Metastable } {/} Tensile strength (MPa) 785 900 6551896 6001795 960970 1024 1033 1030 882975 10991312 10601100 882975 10991312 465950 812 1310 590 1020 1010 7501200 Modulus (GPa) 105 210253 200230 110 105 110 79 75 88113 74-85 75 88113 200 82 100 55 82 66

136

K.S. Katti / Colloids and Surfaces B: Biointerfaces 39 (2004) 133142

The commercial metallic THR implants are ve to six times stiffer than bone and result in signicant problems associated with stress shielding. Ti alloys in the femoral elements of the THR have shown improvement in wear properties. The regenerative and remodeling processes in bone are directly triggered by loading, i.e. bone subjected to loading or stress regenerates and bone not subjected to loading results in atrophy. Thus, the effect of a much stiffer bone implant is to reduce the loading on bone resulting in the phenomenon called as stress shielding. The key problems associated with the use of these metallic femoral stems are thus release of dangerous particles from wear debris, detrimental effect on the bone remodeling process due to stress shielding and also loosening of the implant tissue interface. It has been shown that the degree of stress shielding is directly related to the difference in stiffness of bone and implant material [15,16]. Titanium alloys are favorable materials for orthopedic implants due to their good mechanical properties. However, titanium does not bond directly to bone resulting in loosening of the implant. Undesirable movements at the implant-tissue interface results in failure cracks of the implant. One approach to improving implant lifetime is to coat the metal surface with a bioactive material that can promote the formation and adhesion of hydroxyapatite, the inorganic component of natural bone. The application of bioactive coatings to titanium-based alloys enhance the adhesion of Ti-based implants to the existing bone, resulting in significantly better implant lifetimes than can be achieved with materials in use today. Typically, several silicate glasses are used as bioactive coatings. An ideal bioactive coating would bond tightly both to the bone and the metal. Some ceramic coatings are known to be bioactive and have been tested on Ti implants. However, two problems arise when attempting to coat metals with ceramics. For one, the thermal expansion coefcients of the ceramic and metal are usually different, and as a result, large thermal stresses are generated during processing. These stresses lead to cracks at the interface and compromise coating adhesion. In addition, chemical reactions between the ceramic and metal can weaken the metal in the vicinity of the interface, reducing the strength of the coated system. This problem is particularly important when coating Ti alloys, due to their high reactivity with most oxide materials. Since the modulus of the Ti alloys is lower than that of the CoCrMo alloys, they have been more suitable for THR components. The elastic moduli of the Ti alloys have been engineered to be more suitable by heat treatments resulting in microstructures that have a reduced elastic modulus. The fundamental wear mechanisms of the Ti alloys is still not well understood. Bioglass coatings on Ti implants further improves the biocompatibility of these implants. The glasses are based on mixtures of the oxides of silicon, sodium, potassium, calcium, and magnesium. By adjusting the stoichiometry of the bioglass coating, the thermal expansion coefcient of the glass is made to match that of the Ti alloy, avoiding the generation of thermal stresses.

Also, the glasses become soft at the processing temperature, which is well below the melting point of the Ti alloy. Thus, they ow to uniformly coat the Ti surface. These coatings develop a layer of HAP on their outer surface upon exposure to simulated body uid [17]. Metallic femoral head articulating inside a polymeric (PTFE or UHMWPE) acetabular cup has been one of the most favorable THR element structure [18,19]. Clinical results show that excessive wear and wear debris is the primary cause of failure of UHMWPE or metal implants. Thus, the use of materials with lower modulus and strength such as polymers appear to be more useful for use as bone biomaterials. 2.2. Polymers For orthopedic applications such as xation devices and also use in THR, polymers of very high strength and stiffness are required. The use of polymeric materials in bone biomaterials research is extensive due to many useful properties of polymers. For orthopedic applications, common polymers used are: acrylic, nylon, silicone, polyurethane, ultra high molecular weight polyethylene (UHMWPE), and polypropylene (PP) [20]. Mechanical properties of these polymers are shown in Table 3. Highly stable polymeric systems such as PTFE, UHMWPE or poly(etheretherketone) (PEEK) have been investigated due to their excellent mechanical properties. In the early 1960s, the stainless steel femoral THR component was mated with a PTFE acetabular cup. Poor wearability and distortion in these components prevented further use of PTFE as an important biomaterial for acetabular cups. Acetabular cups made of ultra high molecular weight polyethylene have shown to exhibit superior properties. In the use of acetabular cups made of polyethylene, debris created by wear of polyethylene (PE) articulating surfaces is attacked by the bodys immune system. This leads to bone loss, also known as osteolysis. Since the debris accumulates in the area close to the implant, the bone loss leads to loosening of the implant stem. This results in a repeat surgery. Thus, the main problems associated with the use of PE as acetabular cups is not the wear of the cups
Table 3 Mechanical properties of polymers used in THR [3,23] Material Polymers HDPE UHMWPE PA PS PE PU SR PEEK PTFE PET PMMA UCS (MPa) 25 28 UTS (MPa) 40 21 67 75 35 35 7.6 139 28 61 21 Modulus (GPa) 1.8 1 2.1 2.65 0.88 0.02 0.008 8.3 0.4 2.85 4.5

11.7 144

K.S. Katti / Colloids and Surfaces B: Biointerfaces 39 (2004) 133142

137

themselves but wear of the interfacial adhesion between tissue and implant. Polymers prepared from lactic acid and glycolic acid have been used in the biomedical eld since the 1960s as sutures due to their highly unstable structure leading to bio-degradability [21]. Other biodegradable polymers such as poly(dioxanone) (PDO), poly(trimethylene carbonate) (PTC) copolymers, have been used in the medical eld [22]. As biodegradable polymers, besides PLA and PGA, polycaprolactone (PCL), polyanhydrides (PA), polyorthoesters are also subject of current research. The use of degradable polymers in THR is rather limited due to their inadequate mechanical properties. Due to their degradation properties these polymers have extensive application in tissue engineering. The initial high strength of some degradable polymers such as PLLA has spurred interest in use of these polymers as composite systems with ceramic llers. The use of these materials for composites where stiffening agents are used to enhance mechanical properties is the subject of several current studies. The mechanical properties of polymers used in THR components is shown in Table 3. 2.3. Ceramics As compared to metals, ceramics often cause reduced osteolysis and are regarded as favorable materials for joints or joint surface materials. Several ceramics due to their ease of processing and forming and superior mechanical properties were investigated as bone substitute materials. Conventional ceramics such as alumina were evaluated due to their excellent properties of high strength, good biocompatibility and stability in physiological environments [24]. Due to lack of chemical bonding between sintered alumina and tissue, its applications as a potential bone substitute are limited. Alumina, because of the ability to be polished to a high surface nish and its excellent wear resistance, is often used for wear surfaces in joint replacement prostheses. Femoral heads for hip replacements and wear plates in knee replacements have been fabricated using alumina. In hip replacements, the alumina femoral head is used in conjunction with a metallic femoral stem and an acetabular cup made from UHMWPE for the opposing articulating surface. The wear rates for alumina on UHMWPE have been reported to be as much as 20 times less than that for metal on UHMWPE, making this combination far superior and producing less wear debris. Recently (February 2003) the United States Food and Drug Administration (FDA) has approved alumina ceramic-on-ceramic articulated hips for marketing in the United States. Other ceramic materials have also been investigated for potential applications in orthopedics. The rst paper to report the use of zirconia in biomedical applications was reported in 1969 [25] and the rst paper illustrating the use of zirconia to manufacture ball heads for total hip replacement was reported in 1988 [26]. Considerable research has focused on zirconia and yttria ceramics that are characterized by ne grained microstructures. These ceramics are known as tetragonal zirconia polycrystals (TZP). Zirco-

nia is the material of choice currently for ball heads. Over 300,000 TZP ball heads have been implanted [27] A better match between the bulk material properties of the implant and the bone it replaces can decrease some of the problems associated with using coated metallic implants such as stress shielding. This is often achieved with coatings on implants. Since calcium phosphates are present as apatites in natural bones, researchers have investigated calcium phosphates extensively. Typically, the calcium phosphorus atomic ratios range from 1.5 to 1.67. Tricalcium phosphate (TCP) (Ca3 (PO4 )2 ) and HAP (Ca10 (PO4 )6 (OH)2 are the two minerals at the extremes of this range of calciumphosphorus ratios. Both TCP and HAP are biocompatible materials. Calcium phosphate ceramics, especially HAP and -TCP are widely used for hard tissue replacement due to their biocompatibility and osteoconductive properties [28,29]. As bone defect llers, these ceramics are utilized in powder and block forms. Porous forms with 100300 m pores are preferred since they allow bone to grow into the implant, promoting mechanical xation with the natural bone. The particulate form lacks cohesive strength and lends to dislodge and migrate under externally applied stresses during healing period. In general, the applications of calcium phosphates in the body have been limited by the low strength and low fracture toughness of the synthetic phosphates. Synthetic HAP elicits a direct chemical response at the interface and forms a very tight bond to tissue [30]. Attempts have been made to form high strength consolidated HAP bodies [31,32]. However, its poor mechanical properties such as low strength and limited fatigue resistance restrict its applications. Bending strength as high as 90 MPa has been achieved by colloidal processing of HAP [31]. Mechanical properties of ceramic biomaterials are shown in Table 4. Alumina and titanium dioxide have been used as nanoceramics separately or in nanocomposites with polymers such as polylactic acid or polymethlyl methacrylate. The nanoceramic formulations promote selectively enhanced functions of osteoblasts (bone-forming cells). These functions include cell adhesion, proliferation, and deposition of calcium-containing minerals, an indication of new bone formation in a laboratory setting (Table 5). Ceramics that elicit a favorable bonding to bone tissue are often called as bioactive ceramics. Some compositions
Table 4 Mechanical properties of ceramics used in THR [3,23] Ceramic Zirconia Alumina Bioglass C(Graphite) C(Vitreous) HAP C(LTI pyrolitic) AW glassceramic UCS (MPa) 2000 4000 1000 138 172 600 900 1080 UTS (MPa) 820 300 Modulus (GPa) 220 380 75 25 31 117 28 118

50

138

K.S. Katti / Colloids and Surfaces B: Biointerfaces 39 (2004) 133142

Table 5 Mechanical Properties of composites in comparison to bone Materials Functionally graded: HAP/Yttria, 040% yttria content w/w PHB/HAP, 30% w/w P(HB-co-824% HV)/HAP, 30%w/w P(-hydroxy acids)/HAP Chemically coupled HAP/PE, 740 vol% ller Nano HAP, 3070 to 60 w/w BCP/PLLA, 025% v/v PAAC/HAP PAAC/in situ nano HAP, 4070 w/w PLLA/hydroxyapatite powder, 1030 w/w PLLA/HAP ber, 070% w/w Starch-EVOH (SEVA) blend/HAP, 1030% w/w Starch-EVOH (SEVA) /10% HAP w/w Starch-EVOH (SEVA) /10% HAP w/w with 1% coupling agents (zirconate, titanate and silane) UTS (MPa) Bending strength: 160200 67 6223 18.3420.67 35.878.4 3060 15 2060 Elastic modulus (GPa) 100160 2.52 2.750.47 0.11 0.884.29 (Bending) 2.36.2 2.65 2.255.42 >500 to 2.6 Elongation 12.8 518 26 36.193.2 0.0060.0375 14.70.6% strain 2.44% strain 1.331.99 Elongation at break (%) References [92] [41] [41] [93] [81] [94] [80] [95] [96,97] [76] [83] [44] [45] [45]

42.330.2 53.6 43.349.9

11.8 0.2962.48 (depending on hot pressing parameters) 3.511 1.87.0 3.31 3.754.3

of glasses containing SiO2 , Na2 O, CaO, and P2 O5 bond to soft tissues as well as bone [3335]. The practical use of bioactive glass for THR components has been limited to their use as bioglass coatings on the femoral and acetabular THR components. 2.4. Composites Generally the use of composites for bone biomaterials have included three broad areas: functionally graded composites, polymer-ceramic composites (with and without ber reinforcements), biomimetic composites or composites with biological macromolecules. 2.4.1. Functionally graded composites Composites are fabricated of HAP and zirconia to enhance the mechanical properties of HAP while retaining its bone bonding property. Functionally, graded composites are an important area in composites research. The main feature of a functionally graded composite is the almost continuously graded composition of the composite that results in two different properties at the two ends of the composite. Powder metallurgy methods have been used to make HAP/titanium functionally graded composites offering the biocompatible HAP on the tissue side and titanium for mechanical property [36]. Functionally graded of tricalcium phosphate and uoroapatite composites combine the bioactive properties of uoroapatite with the bioresorbable properties of TCP [37]. The research in this eld is quite promising but currently, the mechanical properties of these composites are clearly in excess of the properties of bone (Table 5).

2.4.2. Polymer-ceramic composites Ceramic polymer composites have superior properties than either ceramics or polymers for use as THR materials [38]. Typically the polymer components have included polymers that have shown good biocompatibility and routinely used in surgical applications. Many polymer composite materials have used HAP as the ceramic ller component [39,40]. Since the polymer materials such as PLA have very low modulus (27 GPa) as compared to that of bone (330 GPa), the HAP needs to be loaded at a very high weight % ratio in the composite. Composites mechanics suggests that a high aspect ratio particle such as a whisker or a ber signicantly improves the modulus with a lower loading wt.%. Thus, the attempts have also been made to prepare needle like or whisker like or brous HAP. Some of these composites such as composites of poly(-hydroxyalkanoates) (PHA) with HAP have shown ultimate strength, elastic modulus and elongation at break similar to bone and are being investigated as potential materials for THR [41]. Calcium carbonate (vaterite) used as a reinforcing agent in poly(lactic acid) composites has shown enhanced mechanical properties such as bending strength of 45 MPa and a modulus as high as 7 GPa with a 050% vat rite loading [42]. Starch-based biodegradable polymers have recently shown potential for applications for bone replacement [43]. Composites based on starch and ethylene vinyl alcohol (EVOH) are known to show degradation when immersed in a simulated body uid. Recently, blends of EVOH (SEVA-C) with starch lled with 1030% by weight of HAP have been fabricated to yield composites with modulus upto about 7 GPa with a 30% HAP loading [44]. Recently, zirconate, titanate and silanes have been used as coupling agents between EVOH and HAP [45]. Optimization of properties with coupling agents is currently an important area of research.

K.S. Katti / Colloids and Surfaces B: Biointerfaces 39 (2004) 133142

139

The bers used for toughening polymeric materials for use in THR also need to be biocompatible. Carbon bers due to their good biocompatibility property have been used to reinforce ultra high molecular weight polyethylene in THR components. Carbon ber-PMMA [46], carbon ber-polypropylene and polysulphone [47,48], carbon ber polyethylene, polybutylene terephthalate, and PEEK [4951] have all been investigated for potential applications for bone plates. The use of these composite materials in THR components has been limited, by the mechanical property mismatch between these composites and the femur bone. Several composite systems such as poly(etheretherketone) PEEK and glass bers [5254] and carbon ber carbon reinforced composites [55,56] have also been investigated as potential bone replacement materials. Multilayered laminated composites of carbon bers and epoxy [57] and braided designs of carbon ber and glass ber epoxy composite [54,58] have been made. Hot pressing mixtures of polymers and HAP bers have also been attempted. HAP bers are fabricated from -Ca(PO3 )2 bers [5961]. Needle like or brous HAP of lengths 1030 m and 0.11 m diameter have been synthesized using hydrothermal synthesis using citric acid [62], 40150 m length and 210 m diameter bers using a solid phase reaction [63]. Bending strength is seen to be almost independent of ber content improvement in modulus from 3.5 to 11 GPa is observed with over 60 wt.% loading of the polymer (polylactic acid) with HAP bers [64]. Although polymeric bers are used in biomedical applications such as absorbable fracture xation systems [65], and scaffolds for tissue engineering [66], the application of polymer bers as a reinforcement phase in THR components is limited due to the inadequate strength and stiffness of the bers. Particulate reinforcement using ceramic phases offers a methodology for improvement in mechanical properties of biomaterials for THR. HAP containing composites retain their useful bioactive properties as well as provide some improvement in mechanical properties. The composites include ber reinforcement of HAP [67,68], HAP/polyethylene [6971], HAP/polyethyl ester [72], HAP/polyphosphasone [73], HAP/polylactide [7476] and HAP/alumina composites [77]. A swelling type biocompatible structural material for bone implants has been investigated recently [78] where the swelling strains are controlled by using a copolymer poly(methyl methacrylate-acrylic acid). In order to improve mechanical properties of such expansion-t materials reinforcement of such copolymers with carbon and Kevlar bers were attempted [56]. Fiber matrix debonding and brillation was observed for Kevlar bers resulting in low modulus and yield strength of Kevlar reinforced composites and a loss in modulus occurred with increasing swelling for the carbon ber reinforced composites. Carbon ber-polysulphone composite has been used for the design of a press-t device for a femoral component of a THR [79]. A self-reinforced polylactide/biphasic calcium phosphate composite has recently been fabricated primarily for use for fracture xation plates [80]. The phosphate content is varied upto 25% by

volume resulting in 515% failure strains and 6030 MPa ultimate tensile stresses. Chemically modied reinforcement phase-matrix interface results in improvement in mechanical properties of composites. Examples of such interface modied composite biomaterials include chemically coupled HAP-polyethylene composites [81], chemically formed HAP-Ca poly(vinyl phosphonate) composites [82] and polylactic acid HAP ber composites [83]. HAP along with bioceramics and bioglasses have been studied extensively as bone repairing material and is used as a coating for implanted prostheses to enhance direct adhesion to bone tissue [84,85]. Bone cements based on PMMA are used to secure orthopedic implants to bone. Due to limited mechanical properties of PMMA, incorporation of HAP in PMMA has been investigated. In addition, enhanced osteogenic properties of the implants is observed with incorporation of HAP in PMMA. [8689]. It has been shown that not only are the mechanical properties of PMMA improved but the osteoblast response of PMMA is also enhanced with addition of HAP [86]. Biosorbable devices made of forged composites of HAP particles and poly l-lactide have shown improved fatigue properties over metallic implants in addition to superior biocompatibility. A new injectable composite for bone repair: poly(-caprolactone) microparticles with biphasic calcium phosphate granules shown some promise [90]. In general the polymer/HAP interfaces are known to have an important role on the resulting mechanical properties [91]. The mechanical properties of various composites investigated in literature for THR materials is shown in Table 4. 2.4.3. Biomimetic composites or composites with biological macromolecules Bone is a nanocomposite of HAP and type I collagen. The HAP-polymer composites are typically simple mixtures fabricated to give a combination of properties of biocompatibility and mechanical strength. Methods to mimic biological processes with synthetic and biological macromolecules has been the focus of recent research. Composites fabricated using co-precipitation of HAP nanocrystals with soluble collagen have been attempted [98100]. Although nanostructure of bone is partially achieved in the HAP/collagen composites, the high cost of type I collagen is an important deterrent in future research in these composites unless less expensive sources of type I collagen are available. HAPgelatin composites are being currently studied for potential bone replacement materials [101]. In addition the biomimetic HAP-embedded collagen nanostructure has inadequate mechanical properties and the proper pore sizes compared to biological bone are not achieved. Attempts are being made in literature to simulate the collagenHAP interfacial behavior in real bone with crosslinking agents such as glutaldehyde [100] with the purpose of potentially improving the mechanical properties of these composites. Other biomimetic routes include in situ mineralization of

140

K.S. Katti / Colloids and Surfaces B: Biointerfaces 39 (2004) 133142

HAP in the presence of polymeric macromolecules such as calcium binding polyacrylic acids of high molecular weight [96,97,102]. Small amounts of these additives during mineralization of HAP have large impact on the bulk mechanical response of the composite [96,97]. The control and development of molecular-level associations of polymer with HAP is suggested to be signicant for the resulting mechanical responses in the composites. Nano HAP has been used to make new in situ composites with polyacrylic acid [102]. The nanostructure of the resulting crystals exhibits a core-shell conguration. In this process, a preferred adsorption of the PAA anions along the surface of developing calcium decient (cd)-HA nanocrystals results. These methods utilize the chemical bonding of nano HAP with PAA to enhance the mechanical properties of the composites. The use of needle-like HAP crystallites in composites may potentially provide new composites with properties more suitable for biomedical applications. In addition, concurrently, composites of biological and synthetic macromolecules with HAP have been extensively studied. Composites of HAP with bone phosphoprotein, bone Gla protein and collagen [103,104] have also been attempted. Mechanical properties in the above composite systems are still inadequate for the potential use of these composites for total bone replacement.

novel material systems where nanostructure and nanointerfaces are included, are key to understanding and predicting the load deformation behavior. With increased computation capabilities such endeavors are becoming increasingly feasible.

Acknowledgements The author acknowledges support from National Science Foundation (CAREER grant #0132768). The cognizant program ofcer is Dr. Larsen-Basse.

References
[1] C.A. Van Blitterswijk, J.J. Grote, W. Kuijpers, W.T. Daems, K.A. de Groot, Biomaterials 7 (1986) 553. [2] J.D. Currey, The Mechanical Adaptations of Bones, Princeton University Press, 1984. [3] F.O. Schmitt, Ann. Rev. Biophys. Biophys. Chem. 4 (1985) 1; M.E. Nimni (Ed.), Collagen: Biotechnology, vol. III, CRC Press, Boca Raton, FL, 1988. [4] K.H. Stenzel, T. Miyata, A.L. Rubin, in: L.J. Mullins (Ed.), Collagen as a Biomaterial in Annual Review of Biophysics and Bioengineering, vol. 3, Annual Reviews Inc., Palo Alto, CA, 1974, p. 231. [5] P. Fratzl, K. Misof, I. Zizak, Fibrillar structure and mechanical properties of collagen, J. Struct. Bio. 122 (1997) 119. [6] D.F. Williams (Ed.), Denitions in Biomaterials: Proceedings of a Consensus Conference of the European Society for Biomaterials, Elsevier, Amsterdam, 1987. [7] J. Black, Biological Performance of Materials: Fundamentals of Biocompatibility, second ed., Marcel Dekker, New York, 1992. [8] Von Recum (Ed.), Handbook of Biomaterials Evaluation, Scientic, Technical and Clinical Testing of Implant Materials, second ed., Taylor and Francis, PA, 1999, p. 915. [9] J. Black, Biological Performance of Materials: Fundamentals of Biocompatibility, Marcel Deckker, New York, 1992. [10] M. Long, H.J. Rack, Biomaterials 19 (1998) 1621. [11] G.K. Mckee, J. Watson-Ferrar, J. Bone Jt. Surg. 48 (1996) 245. [12] P.S. Walker, B.L. Gold, Wear 17 (1971) 285. [13] J.A. Davidson, F.S. Georgette, State-of-the-Art Materials for orthopaedic Prosthetic Devices: on Implant Manufacturing and Material Technology, Proc. Soc. of Manufacturing Engineering, Itasca, IL. [14] John Brunski, in: B.D. Ratner, A.S. Hoffman, F.J. Schoen, J.E. Lemons (Eds.), Biomaterials Science an Introduction to Materials in Medicine, Academic Press, CA, 1996. [15] P. Christel, A. Meunier, A.J.C. Lee (Eds.), Biological and Biomechanical Performance of Biomaterials, Elsevier, Amsterdam, The Netherlands, 1997, pp. 8186. [16] R. Huiskes, E.Y.S. Chao, J. Biomech. 16 (1983) 385. [17] A. Pazo, E. Saiz, A.P. Tomsia, Acta Mater. 46 (1998) 2551. [18] J. Charnley, Lancet (1961) 1129. [19] J. Charnley, Clin. Orthop. Relat. Res. 72 (1970) 7. [20] D.V. Rosato, in: M. Szycher (Ed.) Biocompatible Polymers, Metals and Composites, Technomic Publ., 1983, p. 1022. [21] D.K. Gilding, A.M. Reed, Polymer 20 (1979) 1459. [22] T.H. Barrows, Clin. Mater. 1 (1986) 233. [23] S. Ramakrishna, J. Mayer, E. Wintermantel, K.W. Leong, Comp. Sci. Technol. 61 (2001) 1189. [24] S.F. Hulbert, L.L. Hench, in: P. Vineenzini, High Technology Ceramics, Elsevier, Amsterdam, 1987, pp. 324.

3. Summary and conclusions This paper examines the various materials systems used in design of elements of total hip replacement. The merits and demerits in terms of mechanical properties of these materials systems are also examined. Extensive literature on polymers, ceramics, metals and composites, as next generation potential bone replacement materials exists. In addition several biomimetic methods using biomacromolecules and biomimetic processes are also being studied. Mechanical properties, strength, failure strain and modulus, are either in excess of response for bone or in decit. It has been known that interfaces in these material systems, especially composites play a critical role in observed mechanical properties much like the role of load transfer mechanisms at HAPcollagen interfaces in real bone [105]. It appears that a fundamental molecular understanding of interfacial behavior in these materials, especially composite systems are an area not sufciently addressed in literature. Various experimental characterization techniques using electron microscopy, vibrational spectroscopy, X-ray diffraction, scanning probe microscopy and others are used routinely to characterize these materials besides mechanical property characterization. Finite element modeling of implant systems is also routinely done. It would be benecial to develop robust continuum mechanics analysis techniques with accurate nano-micro and meso structures and interfaces. In addition, atomic scale models for simulating and predicting responses in the

K.S. Katti / Colloids and Surfaces B: Biointerfaces 39 (2004) 133142 [25] J.D. Helmer, T.D. Driskell, Research on bioceramics, in: Proceedings of the Symposium on Use of Ceramics as Surgical Implants, Clemson University, SC, USA, 1969. [26] P. Christel, A. Meunier, J.-M. Dorlot, Ann. N.Y. Acad. Sci. 523 (1988) 234. [27] J. Chevalier, J.M. Drouin, B. Cales, Low temperature ageing behavior of zirconia hip joint heads, in: L. Sedel, C. Rey (Eds.), Bioceramics, vol. 10, Elsevier, Amsterdam, 1977; L.L. Hench, J.W., An introduction to Bioceramics, World Scientic, Singapore, 1993. [28] R.Z. Le Geros, Clin. Mater. 14 (1993) 65. [29] L.L. Hench, J. Wilson, An Introduction to Bioceramics, vol. 1, World Scientic, Singapore, 1993. [30] F.B. Bagambisa, U. Joos, W. Schilli, J. Biomed. Res. 27 (1993) 1047. [31] H.Y. Yasuda, S. Mahara, Y. Umakoshi, S. Imatazo, S. Ebisu, Biomaterials 21 (2001) 2045. [32] L.M. Rodriguez-Lorenzo, M. Valler-Regi, J.M.F. Ferreira, Biomaterials 22 (2001) 583. [33] V. Gross, R. Kinne, H.J. Schmitz, V. Strunz, CRC Critic. Rev. Biocompatibility 4 (1998) 2. [34] L.L. Hench, R.J. Splinter, W.C. Allen, T.K. Greenlec Jr., J. Biomed. Res. Symp. No. 2, Interscience, New York, 1972, p. 117. [35] J. Wilson, G.H. Pigott, F.J. Schoen, L.L. Hench, J. Biomed. Mater. Res. 15 (1981) 805. [36] F. Watari, A. Yokoyama, F. Saso, M. Uo, T. Kawasaki, Compos. Part B: Eng. (UK) 28B (1997) 5. [37] H. Wong, B. Tio, X. Miao, Mater. Sci. Eng. C 20 (2002) 111. [38] M. Wang, S. Deb, K. Tanner, W. Boneld, in: Proceedings of the 7th European Conference on Composite Materials, London, 1996, p. 455. [39] M. Kikuchi, Y. Suetsugu, J. Tanaka, M. Akao, J. Mater Sci. 8 (1997) 361. [40] S. Higashi, T. Yamamuro, T. Nakamura, Y. Ikada, K. Jamshidi, Biomaterials 7 (1986) 183. [41] N. Galego, C. Rozsa, R. Sanchez, J. Fung, A. Vazquez, J.S. Tomas, Polym. Testing 19 (2000) 485. [42] T. Kasuga, H. maeda, K. Kato, M. Nogami, K.-I. Hata, M. Ueda, Biomaterials 24 (2003) 3247. [43] R.L. Reis, A.M. Cunha, J. Mater. Sci. Mater. Med. 6 (1995) 786. [44] R.A. Sousa, J.F. Mano, R.L. Reis, A.M. Cunha, M.J. Bevis, Polym. Eng. Sci. 42 (2002) 1032. [45] C.M. Vaz, R.L. Reis, A.M. Cunha, Biomaterials 23 (2002) 629. [46] S.L.Y. Woo, W.H. Akeson, B. Levenetz, R.D. Coutts, J.V. Mathews, D. Amiel, J. Biomed. Mater. Res. 8 (1974) 321. [47] P.S. Christel, I.L. Leray, L. Sedal, E. Morel, in: G.W. Hastings, D.F. Williams (Eds.), Mechanical Evaluation and Tissue Compatibility of Materials for Composite Bone Plates in Mechanical Properties of Biomaterials, 1980, pp. 367377. [48] L. Claes, W. Hutter, R. Weiss, Mechanical properties of carbon ber reinforced polysulphone plates for internal xation, in: P. Christel, A. Meunier, A.J.C. Lee (Eds.), Biological and Biomechanical Performance of Biomaterials, Elsevier, Amsterdam, The Netherlands, 1997, pp. 8186. [49] N. Ruston, T. Rae, Biomaterials 5 (1984) 352. [50] A. Gillett, S.A. Brown, J.H. Dumbleton, R.P. Pool, Biomaterials 6 (1986) 113. [51] K.A. Jockisch, S.A. Brown, T.W. Bauer, K. Merritt, J. Biomed. Res. 26 (1992) 113. [52] A.A. Corvelli, J.C. Roberts, P.J. Biermann, J.H. Cranmer, J. Mater. Sci. 34 (1999) 2421. [53] M.A. Lopes, F.J. Monteiro, J.D. Santos, Biomaterials 20 (1999) 2085. [54] K. Fujihara, Z.M. Huang, S. Ramakrishna, K. Satknanantham, H. Hamada, Biomaterials 24 (2003) 2661. [55] M. Lewandowska-Szumiel, J. Komender, J. Chlopek, J. Biomed. Mater. Res. 48 (1999) 289.

141

[56] A. Abusaeh, S. Siegler, S.R. Kalidindi, J. Biomed. Mater. Res. (Appl. Biomater.) 38 (1997) 314. [57] F.K. Chang, J.L. Perez, J.A. Davidson, J. Biomed. Mater. Res. 24 (1990) 873. [58] J.A. Simoes, A.T. Marques, G. Jeronimidis, Comp. Sci. Technol. 60 (1999) 559. [59] M. Yoshimura, H. Suda, K. Okamoto, K. Ioku, J. Mater. Sci. 29 (1994) 3399. [60] W. Suchanek, M. Yoshimura, J. Mater. Res. 13 (1998) 94. [61] A. Mourtier, J. Lemaitre, L. Rodvique, Rouxhet, J. Solid State Chem. 78 (1989) 215. [62] M. Yoshimura, H. Suda, K. Okamoto, K. Ioku, J. Mater. Sci. 29 (1994) 3399. [63] Y. Ota, T. Iwashita, T. Kasuga, Y. Abe, J. Am. Ceram. Soc. 81 (1998) 1665. [64] K. Toshihiro, Y. Ota, M. Masayuki, Y. Abe, Biomaterials 22 (2001) 19. [65] M. Vert, P. Christel, H. Garreau, M. Audion, M. Chanavax, F. Chabot, Totally bioresorbable composites systems for internal xation of bone fractures, in: C. Migliaresi, L. Nicolais (Eds.), Polymers in Medicine, vol. 2, Plenum Press, New York, 1986, pp. 263275. [66] C. Vacanti, J. Vacanti, Otolaryngol. Clin. N. Am. 27 (1994) 263. [67] N. Ehsani, A.J. Ruys, C.C. Sorrell, Key Eng. Mater. 104 (1995) 373. [68] A.J. Ruys, K.A. Ziegler A. Brandwood, B.K. Milthorpe, S. Morrey, C.C. Sorrell, Reinforcement of HAP with ceramic and metal bres, in: W. Boneld, G.W. Hastings, K.E. Tanner, (Eds.), Bioceramics, vol. 4, Butterworth-Heinemann, London, 1991, pp. 281286. [69] W. Boneld, Design of bioactive ceramic-polymer composites, in: L.L. Hench, J. Wilson (Eds.), An Introduction to Bioceramics, vol. 1, World Scientic, Singapore, 1993, pp. 290303. [70] M. Wang, D. Porter, W. Boneld, Br. Ceram. Trans. 93 (1994) 91. [71] J. Huang, L. Di Silvio, M. Wang, K. Tanner, W. Bonield, J. Mater. Sci. Mater. Med. 8 (1997) 779. [72] Q. Liu, J. De Wijn, C. Van Blitterssijk, Biomaterials 18 (1997) 1263. [73] C. Reed, K. TenHuisen, P. Brown, H. Allcock, Chem. Mater. 8 (1996) 440. [74] C. Verheyen, C. Klein, J. De Blieck, J. Wolke, C. Van Blitterswijk, K. De Groot, J. Mater. Sci. Mater. Med. 4 (1993) 58. [75] S. Cho, M. Kikuchi, U. Suetsugu, J. Tanaka, Key Eng. Mater. (1997) 132. [76] N. Ignjatovic, S. Tomic, M. Dakic, M.S. Miljkovic, M. Plavsic, D.P. Uskokovic, Biomaterials 20 (1999) 809. [77] J. Li, B. Fartash, L. Hermannsson, Biomaterials 16 (1995) 417. [78] A. Abusaeh, R. Gobran, S.R. Kalidindi, J. Appl. Poly. Sci. 63 (1997) 75. [79] F.P. Magee, A.M. Weinstein, J.A. Longo, J.B. Koeneman, R.A. Yapp, Clin. Orthop. Res. 41 (1988) 235. [80] N.C. Bleach, S.N. Nazhat, K.E. Tanner, M. Kellomaki, P. Tormala, Biomaterials 23 (2002) 1579. [81] M. Wang, W. Boneld, Biomaterials 22 (2001) 1311. [82] Y.E. Greish, P.W. Brown, Biomaterials 22 (2001) 807. [83] T. Kasuga, Y. Ota, M. Nogami, A. Yoshihiro, Biomaterials 22 (2001) 1923. [84] D.C. Tancred, B.A.O. McCormack, A.J. Carr, Biomaterials 19 (1998) 1735. [85] A. Ravaglioli, A. Krajewski, Bioceramics, Chapman and Hall, London, 1992, p. 422. [86] A.M. Moursi, A.V. Winnard, P.L. Winnard, J.L. Lannutti, R.R. Seghi, Biomaterials 23 (2002) 133. [87] C.I. Vallo, P.E. Montemartini, M.A. Fanovich, J.M. Porto Lopez, T.R. Cuadrado, J. Biomed. Res. 48 (1999) 158. [88] Y. Shikinami, M. Okuno, Biomaterials 20 (1999) 859. [89] Y. Shikinami, M. Okuno, Biomaterials 22 (2001) 3197. [90] P. Iooss, A.M. Le Ray, G. Grimandi, G. Daculsi, C. Merle, Biomaterials 22 (2001) 2785.

142

K.S. Katti / Colloids and Surfaces B: Biointerfaces 39 (2004) 133142 [98] M. Kikuchi, Y. Suetsugu, J. Tanaka, S. Ito, S. Ichinose, K. Shiniyama, Y. Hiraoka, Y. Mandia, S. Nakatani, Bioceramics 12 (1999) 393. [99] M.C. Chang, T. Ikonama, M. Kikuchi, J. Tanaka, J. Mater. Sci. Lett. 20 (2001) 1129. [100] M.C. Chang, T. Ikonama, M. Kikuchi, J. Tanaka, J. Mater. Sci. Mater. Med. 13 (2002) 993. [101] M.C. Chang, C.C. Ko, W.H. Douglas, Biomaterials 24 (2003) 2853. [102] S.-C. Liou, S.-Y. Chen, D.-M. Liu, Biomaterials 24 (2003) 3981. [103] D. Bakos, M. Soldan, I. Hernandez-Fuentes, Biomaterials 20 (1999) 191. [104] B.R. Heywood, N.H.C. Sparks, R.P. Shellis, S. Weiner, S. Mann, Connective Tissue Res. 25 (1990) 103. [105] C. Hellmich, F.-J. Ulm, J. Eng. Mech. 123 (2002) 898.

[91] N.L. Ignjatovic, M. Plavsic, M.S. Miljkovic, L.M. Zivkovics, D.P. Uskokovic, J. Microsc. 196 (1999) 23. [92] H. Guo, K.A. Khor, Y.C. Boey, X. Miao, Biomaterials 24 (2003) 667. [93] R. Zhang, P.X. Ma, J. Biomed. Mater. Res. 45 (1999) 285. [94] X. Wang, Y. Li, J. Wei, K. de Groot, Biomaterials 23 (2002) 4787. [95] K. Kato, Y. Eika, Y. Ikada, J. Mater. Sci. 32 (1997) 5533. [96] K. Katti, P. Gujjula, Control of mechanical responses in in situ polymer-hydroxyapatite composites for bone replacement, in: Proceedings of the 15th ASCE Engineering Mechanics Conference, New York, NY, 2002. [97] K.S. Katti, P. Gujjula, A. Ayyarsamy, T. Arens, In situ mineralization of hydroxyapatite for a molecular control of mechanical responses in hydroxyapatite-polymer composites for bone replacement, in: Proceedings of the 2001 MRS Fall Meeting Symposium, vol. GG 4.3.

You might also like