journal of prosthodontic research 59 (2015) 84–95

Available online at www.sciencedirect.com

ScienceDirect journal homepage: www.elsevier.com/locate/jpor

Review

A review of improved fixation methods for dental implants. Part II: Biomechanical integrity at bone–implant interface Yo Shibata DDS, PhDa,*, Yasuhiro Tanimoto PhDb, Noriko Maruyama DDSc, Manamu Nagakura DDSb a

Department of Conservative Dentistry, Division of Biomaterials & Engineering, Showa University School of Dentistry, Tokyo, Japan b Department of Dental Biomaterials, Nihon University School of Dentistry at Matsudo, Chiba, Japan c Department of Orthodontics, Showa University School of Dentistry, Tokyo, Japan

article info

abstract

Article history:

Purpose: The purpose of this article is to review the mechanical requirements of the tissue–

Received 30 October 2014

implant interface and analyze related theories.

Received in revised form

Study selection: The osseointegration capacity of titanium implants has been investigated

1 January 2015

over the past 50 years. We considered the ultimate goal of osseointegration to which form a

Accepted 20 January 2015

desirable interfacial layer and a bone matrix with adequate biomechanical properties.

Available online 19 March 2015

Results: Occasionally, the interface comprises porous titanium and bone ingrowth that enables a functionally graded Young’s modulus, thereby allowing reduction of stress shielding. However,

Keywords:

the optimal biomechanical connection at the interface has not yet been fully clarified. There

Titanium implant

have been publications supporting several universal mechanical testing technologies in terms of

Osseointegration

bone–titanium bonding ability, although the separation of newly formed bone quality is unlikely.

Mechanical properties

Conclusions: The understanding of complex mechanical bone behavior and size-dependent

Finite element analysis

properties ranging from a nano- to a macroscopic level are essential in the biomechanical optimization of implants. The requirements of regenerated tissue at the interface include high strength, fracture toughness related to ductility, and time-dependent energy dissipation and/or elastic–plastic stress distribution. Moreover, a strong relationship between strain signals and peri-implant tissue turnover could be expected, so that ideal implant biomechanics may enable longevity via adaptive bone remodeling. # 2015 Japan Prosthodontic Society. Published by Elsevier Ireland. All rights reserved.

Contents 1. 2. 3.

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Basic mechanical and fatigue properties of titanium . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Basic mechanical properties of bone associated with hierarchical structures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

85 85 86

* Corresponding author at: Department of Conservative Dentistry, Division of Biomaterials & Engineering, Showa University School of Dentistry, 1-5-8 Hatanodai, Shinagawa-ku, Tokyo 142-8555, Japan. Tel.: +81 3 3784 8178; fax: +81 3 3784 8179. E-mail address: [email protected] (Y. Shibata). http://dx.doi.org/10.1016/j.jpor.2015.01.003 1883-1958/# 2015 Japan Prosthodontic Society. Published by Elsevier Ireland. All rights reserved.

85

journal of prosthodontic research 59 (2015) 84–95

4.

5. 6.

1.

Nanoscale mechanical testing technology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.1. Bone mechanical characterization at material level. . . . . . . . . . . . . . . . . 4.2. Mechanical properties of mineralized tissue on titanium surfaces . . . . . 4.3. Molecular structure analysis of mineralized tissue on titanium . . . . . . . 4.4. Bone mechanotransdution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Surface morphology-induced micromechanics and adaptive bone remodeling. Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

Introduction

Orthopedic and dental titanium implants must function as rigid osseous anchors. The biomechanical integrity of implants comprises the mechanical behavior of implant materials, surface-induced bone micromechanics, and adaptive bone remodeling. A diagram of the events presumed to occur at the bone– implant interface is shown in Fig. 1 [1,2]. Once implants are placed with intimate apposition of bone at the surgical site, the immediate response at the interface involves adsorption of tissue fluid and cell binding proteins [1,2]. This critical gap between host bone and the implant surface is likely filled by newly formed bone and a non-collageneous protein-rich cement line [3–5]. Early studies indicated that osteocalcin, osteopontin and bone sialoprotein, as well as certain plasma proteins such as a2HS-glycoprotein, predominate in the cement line [3,4]. Although this protein layer may be responsible for biological bonding to the titanium surface, the inherent mechanical weakness of the protein layer needs to be further stabilized by bone–implant mechanical interlocking [6,7]. Nevertheless, there have been a range of biomechanical issues with titanium implants, such as reduced shear load-bearing properties associated with poor mechanical interlocking at the interface [8,9]. Implant surfaces have been developed by means of several engineering processes, such as grid-blasting, which involves coating the titanium substrate with sintered beads or particles to create a porous layer [10–13], so that the micro– nanoscale external surface textures may address the aforementioned biomechanical issues. Aside from the mechanical interlocking achieved by surface processing parameters, the mechanical requirements of regenerated bone at the interface also need to be addressed. As a consequence of unfavorable osseointegration processes, the implant surface often generates fibrous woven bone and is not replaced by mature lamellar bone [14–17]. Therefore, new biomaterials for bone regenerative purposes, such as titanium implants, need to feature a surface that promotes osteogenic differentiation and proper mineralization during the initial integration stage. The biomechanical integrity of titanium implants has been evaluated based on bone–titanium contact and bone volume fraction by means of histological observation [18,19] and micro-computed tomography [20–22], respectively. However, there has been a mismatch between such observations and the mechanical stability of titanium implants [23,24], allowing us to assume that the mechanical properties of regenerated bone are not entirely associated with observed bone microhistology or densitometry.

. . . . . . . .

. . . . . . . .

. . . . . . . .

. . . . . . . .

. . . . . . . .

. . . . . . . .

. . . . . . . .

. . . . . . . .

. . . . . . . .

. . . . . . . .

. . . . . . . .

. . . . . . . .

. . . . . . . .

. . . . . . . .

. . . . . . . .

. . . . . . . .

. . . . . . . .

. . . . . . . .

. . . . . . . .

. . . . . . . .

. . . . . . . .

. . . . . . . .

. . . . . . . .

. . . . . . . .

. . . . . . . .

. . . . . . . .

. . . . . . . .

. . . . . . . .

. . . . . . . .

. . . . . . . .

87 87 88 88 89 89 91 92

Understanding of basic mechanical properties for titanium and intact bones are indispensable prerequisite so that scientists will begin to adjust the concept of surface modification (characterization) techniques for titanium implants, whether such concepts achieve the biomechanical integrity at bone–implant interface. Bone is a complex hierarchical tissue with different structural levels, namely cortical and trabecular bone at the macroscale, Haversian osteons and lamellae at the microscale, and hydroxyapatite crystals and collagen fibers at the nanoscale [25,26]. The macro–microscale structural variations of bone tissue compromise a precise mechanical evaluation. In this respect, nanoscale mechanical testing technologies enable a measurement of bone mechanical properties at a material level [27] so that more accurate case simulation, according to threedimensional finite element models, is possible. Moreover, an accurate bone remodeling algorithm in the peri-implant region (adaptive bone remodeling) would be obtainable as bone morphology often relates to the stress-strain response associated with the bone mechanical properties at material level [28,29].

2. Basic mechanical and fatigue properties of titanium The mechanical mismatch between host bone and metallic implants has been a longstanding concern. For instance, the elastic modulus of bone is presumed to be 10–30 GPa, while just around 100 GPa for pure titanium, although titanium and its alloys have elastic moduli less than 50% that of cobaltchrome (approximately 230 GPa) [30]. In this context, contacted bone is often inappropriately stress shielded, and hence, implants lose supportive tissue at the peri-implant region over time [31]. Implant elasticity and the long-term bone integrity associated with adaptive bone remodeling are strongly related, as this has been well-established in a range of animal model experiments and clinical trials [32–36]. The mechanical properties of titanium previously reported are summarized in Table 1. Besides its usage in implants, titanium is mainly used as a hard tissue substitute; hence, increased fracture toughness is the basic requirement. In this respect, Ti–6Al–4V alloys have been widely applied as biomedical materials, despite the fact that the toxic element in this alloy is concerning [37]. Meanwhile, b-type titanium alloys are composed of non-toxic elements with a greater strength and toughness balance than that of a + b alloys, such as Ti–6Al–4V [37]. The elastic moduli of b-type titanium alloys are between 55–85 GPa, resulting in elasticity that is much greater than pure titanium and a + b

86

journal of prosthodontic research 59 (2015) 84–95

Fig. 1 – Schematic representation of events at the bone–implant interface. (a) Protein adsorption from blood and tissue fluids, (b) protein desorption, (c) surface changes and material release, (d) inflammatory and connective tissue cells approach the implant, (e) possible targeted release of matrix proteins and selected adsorption of proteins, (f) formation of lamina limitans and adhesion of osteogenic cells, (g) bone deposition on both the exposed bone and implant surfaces, (h) remodeling of newly formed bone [1,2] (license number: 3486160959007). Table 1 – Mechanical properties of titanium alloys for biomedical applications. Alloy 1. Pure Ti grade 1 2. Pure Ti grade 2 3. Pure Ti grade 3 4. Pure Ti grade 4 5. Ti–6Al–4V ELI (mill annealed) 6. Ti–6Al–4V (annealed) 7. Ti–6Al–7Nb 8. Ti–13Nb–13Zr (aged) 9. TMZF (Ti–12Mo–6Zr–2Fe) (annealed) 10. Ti–15Mo (annealed)

Tensile strength (UTS) (MPa)

Yield strength (dy)

Elongation (%)

Modulus (GPa)

Type of alloy

240 345 450 550 860–965 895–930 900–1050 973–1037 1060–1100 874

170 275 450 485 795–875 825–869 880–950 836–908 100–1060 544

24 20 18 15 10–15 6–10 8.1–15 10–16 18–22 21

102.7 102.7 103.4 104.1 101–110 110–114 114 79–84 74–85 78

a a a a a+b a+b a+b b b b

alloys. However, the values are still much higher than that of bone. As homogeneous metallic materials, the mechanical properties of titanium and its alloys, such as tensile characteristics, fracture toughness and fatigue characteristics, have been measured by means of universal mechanical testing machines, in which testing is occasionally carried out in simulated body fluids [37].

3. Basic mechanical properties of bone associated with hierarchical structures Over a long period of human evolution and adaptation, hard tissues such as bone and teeth have developed desirable mechanical properties along with their own hierarchical structures; we can learn useful lessons from such intelligent natural biomaterials. Although the main objective of this review is to address the current biomechanical issues associated with titanium implants, a quick overview of typical natural hard tissue, such as bone, can help us to appreciate some general concepts associated with the excellent design of biomedical materials. Despite the considerable mechanical mismatch between bone and titanium, bone is still functional as a supporting tissue for much harder and stiffer titanium implants, allowing us to conclude that intact bone enables remarkable adaptation associated with strain energy dissipation in the human body. Bone consists of directional structural features across several hierarchical scales, ranging from nanoscale crystals

and molecules to the macroscopic range [38,39]. The foundational unit across the scale is a two-phase arrangement of an anisometric bone mineral (hydroxyl apatite) preferentially oriented in a collagen matrix (Fig. 2) [40,41]. The combination of nanoscale apatite crystals and collagen fibers enables light but tough mechanical properties [42,43]. Based on Griffith’s theory of brittle fracture [44], the smaller the individual sample, the higher the strength is, because smaller samples contain less and smaller defects. In this respect, nanoscale bone apatite crystals enable a higher strength than a bulk sintered apatite block. However, understanding bone mechanical properties is quite complicated because of the size-dependent mechanical properties associated with a heterogeneous, anisotropic material with complex, multiscale structural variations [45]. As a consequence of the multiscale structure, mechanical properties, such as the elastic modulus of bone, have varied dramatically in accordance with differently scaled mechanical testing technologies. The elastic modulus of large tensile cortical bone has been shown to be approximately 14–20 GPa, while that attained from micro-bending tests was 5 GPa [38]. Additionally, the effects of specimen size on the observed moduli suggest that the micro-structural constitution of bone may not be constant, similar to macroscopic observation. As mentioned above, the basic building block of bone, comprising apatite crystals and collagen fibers, is extremely small. The precise mechanical evaluation at a material level by means of universal mechanical testing technology is therefore nearly impossible.

87

journal of prosthodontic research 59 (2015) 84–95

Fig. 2 – Schematic illustration of hierarchical structure of bone (license number: 3495260148887).

Load, P

(hmax , Pmax) S

P = Pmax Original Deformed

θ

hc

hmax 0 0

hc displacement, h

Fig. 3 – Nanoindentation force displacement curve for measurement of hardness and elastic modulus.

4.

Nanoscale mechanical testing technology

Given the above concerns, nanoscale mechanical testing technologies, such as nanoindentation, may resolve specific features of bone mechanical properties through precise measurement protocols [27]. The instrumented indentation technique has become a popular method to measure the mechanical properties of different materials at the same measurement scale, so that even multiscale structures of bone or teeth can be comparable to artificial biomaterials at the material level. Indeed, nanoindentation mechanical testing has been introduced to evaluate the intrinsic mechanical properties of regenerated bone within a titanium bone chamber, so that the mechanical quality of bone on machined and roughened titanium implants can be predicted [46]. Most nanoindentation experiments use a sharp pyramidal Berkovich indenter from which the hardness and elastic modulus can be determined based on the compliance methods advocated by Oliver and Pharr [47,48]. Briefly, the indenter tip contacts the material surface with an applied loading force followed by unloading, so that a transducer can visualize a load–displacement curve (Fig. 3). Nanoindentation uses the ‘‘area function’’ based on a certain tip geometry [48]. The area function provides a method of gaining an accurate contact area (A) against the indenter tip penetration depth. The onset of the unloading slope reveals the contact stiffness of the

material’s surface. The hardness (H) and reduced (elastic) modulus (Er) can be calculated according to the equations below, where S is the contact stiffness calculated as the slope of the unloading curve at the onset of unloading, P is the applied loading force, and A is the projected area of the indenter tip as a function of the contact depth: H ¼ P=A

(1)

pffiffiffi s p Er ¼ pffiffiffiffi 2 A

(2)

4.1.

Bone mechanical characterization at material level

The nanoindentation compliance method has been developed on the basis of a presumption that the indenter is always contacting a pure elastic–plastic, time-independent material such as ceramic or metallic materials [47,48]. However, biological tissue often generates time-dependent viscoelastic behavior, similar to polymeric artificial materials [49–51]. Although time-dependent behavior is presumed to be a function of bone durability [52–54], the delayed responses compromise the precise measurement of bone mechanical properties. According to a simple viscoelastic Maxwell model represented by a viscous dashpot and elastic spring connected in series [55], deformation of viscoelastic materials may be associated with strain rates. For instance, the viscous dashpot

journal of prosthodontic research 59 (2015) 84–95

Thold Pmax

PU

Load, P

Load, P

88

S Se

PL

0 0

Time, t

0 0

Displacement, h

Fig. 4 – Viscoelastic hysteresis loop forms during unloading.

cannot instantaneously respond to high strain rates, thereby resulting in less deformation and enhanced apparent stiffness. Alternatively, the most readily observed delayed (timedependent) response from a material surface is creep behavior during constant holding indentation, as the viscous dashpot is largely in motion against such lower strain rates [27]. Consequently, a hysteresis loop forms during unloading, upon which loading by creep strain rates generally results in overestimation of contact stiffness (Fig. 4) and the projected contact area, and hence, the observed mechanical properties such as hardness and elastic moduli are varied even in nanoindentation experiments [56,57]. Despite the substantial concerns associated with time-dependent properties, a number of bone nanoindentation experiments have already been carried out. That is to say, most researchers rule out the inaccuracy of complex elastic moduli against different strain rates due to the viscoelastic effects. Separation of elastic (storage) modulus and viscoelasticity (loss modulus) through either quasi-static or dynamic mechanical analysing has been reported [45,56,58–60]; namely, these offsets would be highly appreciated with respect to measuring bone nanomechanical properties.

4.2. Mechanical properties of mineralized tissue on titanium surfaces There are still few studies determining the biomechanical properties of the mineralized layer on titanium implants, and more studies are essential. In vitro bone cell behavior on a titanium sample can be similar to that of an in vivo afibrillar mineralized layer, and cell organization is comparable to that in a mineralized matrix on a titanium surface during culture with supplemental ascorbic acid, sodium b-glycerophosphate, and dexamethasone [61]. This is therefore an important consideration as the information obtained reflects in vivo events. The biomechanical evaluation of such micro-environments is challenging. Nano-biomechanical testing technology, such as nanoindentation, enables measurements of the mechanical properties of samples of very small volume [62– 64]. There are two crucial factors associated with the evaluation of in vitro mineralized tissue on titanium surfaces. First—and this is central to standard nanoindentation analysis—is the assumption that the surface is in ideal contact with the indenter penetration depth. In reality, all biological surfaces are characterized by some degree of roughness or heterogeneity [65]. Secondly, the

viscoelastic properties of biological tissues can cause problems [57]. In elastic–plastic materials, the obtained values are always constant, regardless of the load function. However, many polymeric biomaterials and biological tissues exhibit timedependent or viscoelastic behavior (Section 4.1) [27]. The observed effect of viscoelasticity on indentation is creep, or a sinking of the indenter tip into the sample under a constant load function. In such viscoelastic materials, the hardness and moduli usually vary with load function (or loading rate) because different loadings affect the creep or sinking behavior, leading to variations in the contact area (A) and unloading stiffness (S) in Eq. (2). A depth-dependent loading/partial unloading function consists of a number of loading/unloading portions (data points) toward the applied final loading force (Fig. 5). A partial unloading technique enables ideal contact between the indenter tip and the sample surface as a function of depth, so the depth-dependent mechanical properties of heterogeneous biological structures can be normalized in a direction perpendicular to the sample surface [66–68]. An additional holding time at each peak load, followed by an unloading portion, prevents an inappropriate unloading slope because the hold enables relaxation of the initial loading force. Numerous mineralized tissues on titanium surfaces can therefore be compared using average depth-dependent nanomechanical values. The mechanical properties of mineralized tissue are associated with maturation of the extracellular mineralized bone matrix proteins produced by osteoblasts. The increased hardness and elastic modulus of mineralized tissue is only observed on titanium surfaces with the release of hydroxyl radicals [55,66,67]. In the living body, oxidation of lysine residues in immature bone matrix proteins, such as collagen molecules, is the final enzyme reaction for cross-linking between molecules [69,70]. Therefore, the reduction of oxidation enzymes generates inferior mechanical properties of in vitro mineralized tissues, even if osteogenic differentiation is greatly increased by a supplemental anabolic agent such as bone morphogenetic protein-2 [68]. Aside from the natural oxidation enzymes, surface hydroxyl radicals may enable oxidation of the lysine residue from immature collagen molecules. Surface radical reactions on titanium may play an important role in the biomechanical considerations at the interface between healing bone and titanium implants.

4.3. Molecular structure analysis of mineralized tissue on titanium Raman spectroscopy analyses and alterations in the protein structure are of particular importance in distinguishing mature or immature calcified tissues, since subtle molecular changes often cause detectable vibrational changes [71–74]. The coupling of a Raman spectrometer with an optical microscope allows reduction of the analysis scale to the micrometer scale, with better spatial resolution than that obtained using infrared spectroscopy. There have been several characteristics associated with Raman spectra of calcified

89

Force (μN)

journal of prosthodontic research 59 (2015) 84–95

0

20

40

60

80

100

Time (sec)

Elastic modulus (GPa)

25 20 15 10 5 0 0

20

40

60

80 100 120 Contact depth (nm)

140

160

180

200

Fig. 5 – A depth-dependent load/partial unloading load function and consequent effective measurement range on cortical bone.

tissues (Fig. 6). Briefly, the ‘Amide I’ region between 1600 and 1700 cm1 revealed the occurrence of three or more bands at 1623, 1637 and 1656 cm1, attributable to characteristic collagen secondary structures [72,73]. The band at 1637 cm1 is characteristic of a collagen secondary structure, and is evident as a shoulder at the ‘Amide I’ band in the Raman spectrum of collagen. The increase in the ‘Amide I’ band shoulder at 1656 cm1 was attributable to the maturity of typical a-helix collagen secondary structures. This is related to hydrogen bonding between collagen molecules and subsequent cross-linking of tissues [75]. The Raman spectra also show bone mineral bands from apatite phosphate at 960 cm1 and carbonate at 1070 cm1. The ratio of mineral bands in the ‘Amide III’ band at around 1250 cm1, attributable to disordered and ordered bone matrix proteins, indicates the mineral to matrix ratio of the calcified tissues [74]. An inorganic apatite transformation associated with collagen cross-linking plays an important role in the biological stability of the mineralized layer at the interface between bone and titanium implants, thereby increasing the longevity of the implant prosthesis.

4.4.

lining cells for the sensing of mechanical stimuli [77]. Bone is often considered to be a fluid saturated porous network [60]. The application of mechanical strains causes an imbalance of pore pressures so that bone fluid pressure generates fluid flow. Such interstitial fluid flow is the most potential way of informing bone cells with respect to mechanical loading [78–80]. A bone nanoindentation study revealed a pile-up response against high-strain indenter penetration because of reduced fluid permeability within the bone which with time would diffuse radially and to a reduction in pile-up during constant load or displacement [60]. Along with the recovery process, proteins refolding and fluid permeability or flow back into the deformed area would occur. The large fluid flow as a consequence of pile-up response was only observed in highly mineralized intact bone while the higher collagen and fluid (porosity) content of lower mineralized bone is less able to generate a pile-up response and consequent fluid motion. Continuous balanced osseous remodeling is essential for the successful maintenance of implants. Thus, mechanotransduction due to the proper peri-implant bone tissue strains (mechanical properties) is an important determinant for prolonged biomechanical stability of titanium implant.

Bone mechanotransdution

Bone remodeling is dependent on cellular processes, namely bone formation by osteoblasts and bone resorption by osteoclasts. The detection of the applied stress is done by certain sensor cells such as bone lining cells of osteoblastic origin and osteocytes[76]. The cells transduce obtained mechanical signal and modulate bone formation and resorption (see also Section 5). The osteocytes are differentiated osteoblasts. A number of researchers have focused on the role of osteocytes in the mechanotransduction process [76]. The osteocytes form noncalcified three-dimensional network connected to the bone

5. Surface morphology-induced micromechanics and adaptive bone remodeling The mechanical stability of titanium implants has been primarily achieved by an intimate apposition in the bone surgical site and subsequent bone ingrowth to the implant surface. With increased demand for implants in orthopedic and oral/maxillofacial surgery, rough titanium surfaces have proven rather effective in improving osteogenic differentiation and micro-scale biomechanics in parallel [81,82]. Meanwhile, biomechanical evaluations of dental implants using

90

1800

1600

1400

1200

1000

800

652 633

732

793

875

958

1047 1039

1175 1156

1321 1263

1350

1479 1460 1443

1541

1620 1575

Raman intensity (a.u.)

1249

journal of prosthodontic research 59 (2015) 84–95

600

Raman shifts (cm-1)

Assignment

732

DNA

793

DNA

875

Hydroxyproline

~958

Apatite phosphate

~1070

Carbonate

1249

Collagen, amide III

1263

Collagen, amide III

1350

Glycosaminoglycans

1440~

Collagen-other proteins

1575

DNA

1600~

Collagen, amide I

Raman intensity (a.u.)

Raman shift (cm-1)

Raman shift (cm-1)

Fig. 6 – Peak assignments on mineralized tissue. the ‘Amide I’ region between 1600 and 1700 cmS1 revealed the occurrence of three or more bands at 1623, 1637 and 1656 cmS1, attributable to characteristic collagen secondary structures. The band at 1637 cmS1 is characteristic of a collagen secondary structure, and is evident as a shoulder at the ‘Amide I’ band in the Raman spectrum of collagen. The increase in the ‘Amide I’ band shoulder at 1656 cmS1 was attributable to the maturity of typical a-helix collagen secondary structures. This is related to hydrogen bonding between collagen molecules and subsequent cross-linking of tissues.

finite element analysis (FEA) have been widely performed by many researchers. As is well-known, FEA is a powerful tool for analysing stress distribution in peri-implant tissues, and most of the corresponding mechanical properties used in the FEA model, such as Young’s modulus and Poisson’s ratio of titanium, cortical and cancellous bone, are often obtained from relevant literature. Accordingly, material properties used in FEA by several studies [83–90] are shown in Table 2. By means of FEA, the effect of surface roughness treatment such as polishing, sandblasting, and plasma-spraying (or porousbeading) on the distribution of stresses at the bone–implant interface in immediately-loaded mandibular implants has been investigated [91,92]. Of these, surface treatments were represented by the coefficient of friction. More recently, enhanced mechanical interlocking has been advocated by bone tissue growth through porous titanium coated surfaces [30,93,94]. Upon titanium porosity, elastic properties closer to those of bone than to solid titanium surfaces enable reduction

of stress shielding in peri-implant tissues. In FEA, poroussurfaced implants appeared to distribute stress in a more uniform pattern around the implant when compared with smooth-surfaced implants [95]. It was indicated by FEA that the low modulus implant induces a stress distribution closer to the actual physiological phenomenon [96]. In addition to these features, the pores on titanium implants increase the surface area for osseous apposition [81,82]. Accordingly, porous coating on implants is of particular interest and has been given widespread attention by researchers as a method to reduce the stiffness mismatches between bone and titanium, achieving stable longevity of titanium implants by means of functionally graded biomechanics at the interface [30]. Aside from the load-bearing properties, a previous radiography study to examine the bone remodeling of porous-coated implants in dogs found that the implant with greater porosity of the coated area stabilized more quickly than that with a less

91

journal of prosthodontic research 59 (2015) 84–95

Table 2 – Biomechanical evaluations of dental implant–homogeneous bone models by using FEA. Material properties useda

Author (year)

Significance

Titanium implant

Cortical bone

Cancellous bone

Harirforoush et al. (2014)

E = 104.8 GPa; n = 0.34; r = 4430 kg/m 3

E = 14 GPa; n = 0.3; r = 1300 kg/m 3

E = 0.49 GPa; n = 0.3; r = 1300 kg/m 3

Pen˜arrocha et al. (2013)

E = 117 GPa; n = 0.3

E = 13 GPa; n = 0.3

E = 1.37 GPa; n = 0.3

Ferraz et al. (2012)

E = 110 GPa; n = 0.35

E = 1.4 GPa; n = 0.3

E = 1.37 GPa; n = 0.3

Ausiello et al. (2012)

E = 120 GPa; n = 0.33

E = 13.7 GPa; n = 0.30

E = 0.50 GPa; n = 0.30

Ding et al. (2009)

E = 103.4 GPa; n = 0.35

E = 13.7 GPa; n = 0.30

E = 1.37 GPa; n = 0.30

Baggi et al. (2008)

E = 114 GPa; n = 0.34

E = 13.7 GPa; n = 0.3

E = 0.50 GPa; n = 0.3

Tanimoto et al. (2006)

E = 107 GPa; n = 0.34; r = 4500 kg/m3; h = 0.0003 E = 107 GPa; n = 0.33

E = 13.7 GPa; n = 0.30; r = 2000 kg/m3; h = 0.0100 E = 13.7 GPa; n = 0.30

E = 0.69 GPa; n = 0.30; r = 1000 kg/m3; h = 0.0381 E = 1.37 GPa; n = 0.30

Kitagawa et al. (2005) a

The resonace frequency of the implant was strongly influenced by the cortical bone to cancellous bone ratio at the implant/bone interface. The implant neck design and implantabutment joint types influence peri-implant bone stresses and abutment micromovement. The implant with microthreads showed higher stress concentration for cortical bone compared with the smooth implant, and lower stress concentration for cancellous bone. The implant thread design plays a crucial role to reduce induced stresses and damage in bone. The stress and strain on the implant-bone interfaces significantly decreased with increasing the implant diameter. The implant design (i.e., diameter and length), crestal bone geometry, and placement site affect the mechanisms of load transmission. The implant with stress-absorbing element had a higher damping effect than the implant without stress-absorbing element. The stress in bone was sensitive to the thickness and elastic modulus of cortical bone.

References

[87]

[89]

[86]

[83]

[85]

[84]

[90]

[88]

E: elastic modulus; n: Poisson’s ratio; r: density; h: loss factor.

porous coated area [97]. Thus, implant osseointegration is considered to involve a combination of initial bone ingrowth onto the surface and adaptive bone remodeling against strain energy density over time [28,29]. Bone is most likely to have the ability to adapt against strain energy. Several attempts to quantify the adaptive bone remodeling process have been reported in the literature based on the widely accepted original hypothesis of Wolff’s theory [98–101]. These studies implied that bone has internal strain sensors measuring the internal loading condition and transducers releasing the signals, so that bone remodeling is activated by a cooperative interaction between osteogenic and osteoclastic cells. The internal mechanical loading such as strain energy density in bone micro-structures may be determined mathematically by FEA [28,102–104]. An example by Weinans et al. suggested resemblance to the density distribution of a proximal human femur by means of a twodimensional FEA model [28]. Given the fact that intact bone is likely to have a selfoptimization capability, the biomechanical integrity of the implant–bone interface is a critical indicator for implant stability and longevity. Rungsiyakull et al. conducted surface morphology (porosity) optimization for titanium implants in terms of interfacial biomechanics and adaptive bone remodeling in the peri-implant region using a FEA model with a multiscale remodeling algorithm [29]. The macroscopic model generates the global responses of bone, while the microscopic level quantifies the change in the apparent

bone density of the representative volume elements. The simulation of adaptive bone remodeling in the study by Rungsiyakull et al. suggests that increasing the volume fraction of the coating particles would result in increased bone density over time, as well as achieving a better initial stability than those with a lower volume fraction [29]. As shown above in Table 1, homogeneous models for bone structure have been employed extensively in FEA studies. This is mainly because linear analysis is useful for analysing the stress distribution of homogeneous materials; however, bone is hierarchical in structure and induces timedependent bone remodeling. Accordingly, bone mechanical properties should be considered on the basis of an empirical relationship between elastic modulus and bone mineral density [105–107]. In this respect, biomechanical simulation of titanium implants should be performed with caution given to the aforementioned difficulty in terms of complex bone mechanical properties related to time-dependency [108,109].

6.

Conclusion

Although the titanium implant surface enables a biological bonding capability through the cement layer comprising body fluid or proteins, the weaknesses of the protein layer need to be further stabilized by microscale mechanical interlocking. Roughened or porous implant surfaces should accommodate

92

journal of prosthodontic research 59 (2015) 84–95

enhanced biomechanical properties due to the maximized contact area between newly formed bone and the implant surface. Understanding titanium and basic bone mechanical properties is indispensable prerequisite for the development and clinical use of implants. Bone mechanical properties are quite complicated and are still not fully understood. Thus, precise prediction of implant biomechanics has long been a challenge. Nanomechanical testing technology is a useful tool for achieving greater precision with respect to the biomechanical response, including adaptive bone remodeling in the periimplant region.

Conflict of interest The authors have no conflicts of interest to declare.

references

[1] Puleo DA, Nanci A. Understanding and controlling the bone–implant interface. Biomaterials 1999;20:2311–21. [2] Shibata Y, Hosaka M, Kawai H, Miyazaki T. Glow discharge plasma treatment of titanium plates enhances adhesion of osteoblast-like cells to the plates through the integrinmediated mechanism. Int J Oral Maxillofac Implants 2002;17:771–7. [3] Ayukawa Y, Takeshita F, Inoue T, Yoshinari M, Shimono M, Suetsugu T, et al. An immunoelectron microscopic localization of noncollagenous bone proteins (osteocalcin and osteopontin) at the bone–titanium interface of rat tibiae. J Biomed Mater Res 1998;41:111–9. [4] Linder L. High-resolution microscopy of the implant– tissue interface. Acta Orthop Scand 1985;56:269–72. [5] Murai K, Takeshita F, Ayukawa Y, Kiyoshima T, Suetsugu T, Tanaka T. Light and electron microscopic studies of bone-titanium interface in the tibiae of young and mature rats. J Biomed Mater Res 1996;30:523–33. [6] Brunski JB. In vivo bone response to biomechanical loading at the bone/dental–implant interface. Adv Dent Res 1999;13:99–119. [7] McKee MD, Nanci A. Osteopontin at mineralized tissue interfaces in bone, teeth, and osseointegrated implants: ultrastructural distribution and implications for mineralized tissue formation, turnover, and repair. Microsc Res Tech 1996;33:141–64. [8] Kutty MG, Bhaduri S, Bhaduri SB. Gradient surface porosity in titanium dental implants: relation between processing parameters and microstructure. J Mater Sci Mater Med 2004;15:145–50. [9] Rosa AL, Crippa GE, De Oliveira PT, Taba Jr M, Lefebvre LP, Beloti MM. Human alveolar bone cell proliferation, expression of osteoblastic phenotype, and matrix mineralization on porous titanium produced by powder metallurgy. Clin Oral Implants Res 2009;20: 472–81. [10] Bobyn JD, Pilliar RM, Cameron HU, Weatherly GC. The optimum pore size for the fixation of porous surfaced metal implants by the ingrowth of bone. Clin Orthop 1980;150:263–70. [11] Pilliar RM. Porous-surfaced metallic implants for orthopedic applications. J Biomed Mater Res 1987;21:1–33. [12] Thelen S, Barthelat F, Brinson LC. Mechanics considerations for microporous titanium as an orthopedic implant material. J Biomed Mater Res A 2004;69:601–10.

[13] Wen CE, Yamada Y, Shimojima K, Chino Y, Asahina T, Mabuchi M. Processing and mechanical properties of autogenous titanium implant materials. J Mater Sci Mater Med 2002;13:397–401. [14] Becker W, Schenk R, Higuchi K, Lekholm U, Becker BE. Variations in bone regeneration adjacent to implants augmented with barrier membranes alone or with demineralized freeze-dried bone or autologous grafts: a study in dogs. Int J Oral Maxillofac Implants 1995;10:143–54. [15] Berglundh T, Abrahamsson I, Lang NP, Lindhe J. De novo alveolar bone formation adjacent to endosseous implants: a model study in the dog. Clin Oral Implants Res 2003;14:251–62. [16] Deguchi T, Takano-Yamamoto T, Kanomi R, Hartsfield Jr JK, Roberts WE, Garetto LP. The use of small titanium screws for orthodontic anchorage. J Dent Res 2003;82: 377–81. [17] Roberts WE, Smith RK, Zilberman Y, Mozsary PG, Smith RS. Osseous adaptation to continuous loading of rigid endosseous implants. Am J Orthod 1984;86:95–111. [18] Isidor F. Histological evaluation of peri-implant bone at implants subjected to occlusal overload or plaque accumulation. Clin Oral Implants Res 1997;8:1–9. [19] Paolantonio M, Dolci M, Scarano A, d’Archivio D, Di Placido G, Tumini V, et al. Immediate implantation in fresh extraction sockets. A controlled clinical and histological study in man. J Periodontol 2001;72:1560–71. [20] Bernhardt R, Scharnweber D, Mu¨ller B, Thurner P, Schliephake H, Wyss P, et al. Comparison of microfocusand synchrotron X-ray tomography for the analysis of osteointegration around TI6AL4V-implants. Euro Cell Mater 2004;7:42–51. [21] Bernhardt R, Van Den Dolder J, Bierbaum S, Beutner R, Scharnweber D, Jansen J, et al. Osteoconductive modifications of Ti-implants in a goat defect model: characterization of bone growth with SR mCT and histology. Biomaterials 2005;26:3009–19. [22] 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: threedimensional micro-CT based structural analyses of porous bioactive titanium implants. Biomaterials 2006;27:5892–900. [23] Gao Y, Luo E, Hu J, Xue J, Zhu S, Li J. Effect of combined local treatment with zoledronic acid and basic fibroblast growth factor on implant fixation in ovariectomized rats. Bone 2009;44:225–32. [24] Liu S, Broucek J, Virdi AS, Sumner DR. Limitations of using micro-computed tomography to predict bone–implant contact and mechanical fixation. J Microsc 2012;245:34–42. [25] Ionova-Martin SS, Do SH, Barth HD, Szadkowska M, Porter AE, Ager Iii JW, et al. Reduced size-independent mechanical properties of cortical bone in high-fat dietinduced obesity. Bone 2010;46:217–25. [26] Chang YT, Chen CM, Tu MY, Chen HL, Chang SY, Tsai TC, et al. Effects of osteoporosis and nutrition supplements on structures and nanomechanical properties of bone tissue. J Mech Behav Biomed Mater 2011;4:1412–20. [27] Ebenstein DM, Pruitt LA. Nanoindentation of biological materials. Nano Today 2006;1:26–33. [28] Weinans H, Huiskes R, Grootenboer HJ. The behavior of adaptive bone-remodeling simulation models. J Biomech 1992;25:1425–41. [29] Rungsiyakull C, Li Q, Sun G, Li W, Swain MV. Surface morphology optimization for osseointegration of coated implants. Biomaterials 2010;31:7196–204. [30] Ryan G, Pandit A, Apatsidis DP. Fabrication methods of porous metals for use in orthopaedic applications. Biomaterials 2006;27:2651–70.

journal of prosthodontic research 59 (2015) 84–95

[31] Kro¨ger H, Venesmaa P, Jurvelin J, Miettinen H, Suomalainen O, Alhava E. Bone density at the proximal femur after total hip arthroplasty. Clin Orthop 1998;6:6–74. [32] Pilliar RM, Cameron HU, Binnington AG, Szivek J, Macnab I. Bone ingrowth and stress shielding with a porous surface coated fracture fixation plate. J Biomed Mater Res 1979;13:799–810. [33] Engh CA, Bobyn JD. Principles, techniques, results, and complications with a porous-coated sintered metal system. Instr Course Lect 1986;35:169–83. [34] Bobyn JD, Glassman AH, Goto H, Krygier JJ, Miller JE, Brooks CE. The effect of stem stiffness on femoral bone resorption after canine porous-coated total hip arthroplasty. Clin Orthop 1990;19:6–213. [35] Bobyn JD, Mortimer ES, Glassman AH, Engh CA, Miller JE, Brooks CE. Producing and avoiding stress shielding: laboratory and clinical observations of noncemented total hip arthroplasty. Clin Orthop 1992;7:9–96. [36] Sychterz CJ, Timmie Topoleski LD, Sacco M, Engh Sr CA. Effect of femoral stiffness on bone remodeling after uncemented arthroplasty. Clin Orthop 2001;218–27. [37] Niinomi M. Mechanical properties of biomedical titanium alloys. Mater Sci Eng A 1998;243:231–6. [38] Rho JY, Kuhn-Spearing L, Zioupos P. Mechanical properties and the hierarchical structure of bone. Med Eng Phys 1998;20:92–102. [39] Weiner S, Wagner HD. The material bone: structure– mechanical function relations. Annu Rev Mater Sci 1998;28:271–98. [40] Asefa T, Yoshina-Ishii C, MacLachlan MJ, Ozin GA. New nanocomposites: putting organic function ‘inside’ the channel walls of periodic mesoporous silica. J Mater Chem 2000;10:1751–5. [41] Zimmermann EA, Gludovatz B, Schaible E, Busse B, Ritchie RO. Fracture resistance of human cortical bone across multiple length-scales at physiological strain rates. Biomaterials 2014;35:5472–81. [42] Wagner HD, Weiner S. On the relationship between the microstructure of bone and its mechanical stiffness. J Biomech 1992;25:1311–20. [43] Balooch M, Habelitz S, Kinney JH, Marshall SJ, Marshall GW. Mechanical properties of mineralized collagen fibrils as influenced by demineralization. J Struct Biol 2008;162:404–10. [44] Griffits AA. The phenomena of rupture and flow in solids. Masinovedenie 1995;9–14. [45] Isaksson H, Nagao S, Małkiewicz M, Julkunen P, Nowak R, Jurvelin JS. Precision of nanoindentation protocols for measurement of viscoelasticity in cortical and trabecular bone. J Biomech 2010;43:2410–7. [46] Butz F, Aita H, Wang CJ, Ogawa T. Harder and stiffer bone osseointegrated to roughened titanium. J Dent Res 2006;85:560–5. [47] Pharr GM, Oliver WC, Brotzen FR. On the generality of the relationship among contact stiffness, contact area, and elastic modulus during indentation. J Mater Res 1992;7:613–7. [48] Oliver WC, Pharr GM. Improved technique for determining hardness and elastic modulus using load and displacement sensing indentation experiments. J Mater Res 1992;7:1564–80. [49] Kjær M. Role of extracellular matrix in adaptation of tendon and skeletal muscle to mechanical loading. Physiol Rev 2004;84:649–98. [50] Ji B, Gao H. Mechanical properties of nanostructure of biological materials. J Mech Phys Solids 2004;52:1963–90. [51] Cristofolini L, Viceconti M, Cappello A, Toni A. Mechanical validation of whole bone composite femur models. J Biomech 1996;29:525–35.

93

[52] Bembey AK, Bushby AJ, Boyde A, Ferguson VL, Oyen ML. Hydration effects on the micro-mechanical properties of bone. J Mater Res 2006;21:1962–8. [53] Huang G, Lu H. Measurements of two independent viscoelastic functions by nanoindentation. Exp Mech 2007;47:87–98. [54] Oyen ML. Poroelastic nanoindentation responses of hydrated bone. J Mater Res 2008;23:1307–14. [55] Wenchang T, Wenxiao P, Mingyu X. A note on unsteady flows of a viscoelastic fluid with the fractional Maxwell model between two parallel plates. Int J Non-Linear Mech 2003;38:645–50. [56] Fujisawa N, Swain MV. Nanoindentation-derived elastic modulus of an amorphous polymer and its sensitivity to load-hold period and unloading strain rate. J Mater Res 2008;23:637–41. [57] Shibata Y, He LH, Kataoka Y, Miyazaki T, Swain MV. Micromechanical property recovery of human carious dentin achieved with colloidal nano-b-tricalcium phosphate. J Dent Res 2008;87:233–7. [58] Cheng YT, Ni W, Cheng CM. Determining the instantaneous modulus of viscoelastic solids using instrumented indentation measurements. J Mater Res 2005;20:3061–71. [59] Feng G, Ngan AHW. Effects of creep and thermal drift on modulus measurement using depth-sensing indentation. J Mater Res 2002;17:660–8. [60] Maruyama N, Shibata Y, Wurihan A, Swain MV, Kataoka Y, Takiguchi Y, et al. Strain-rate stiffening of cortical bone: observations and implications from nanoindentation experiments. Nanoscale 2014;6:14863–71. [61] Irie K, Zalzal S, Ozawa H, McKee MD, Nanci A. Morphological and immunocytochemical characterization of primary osteogenic cell cultures derived from fetal rat cranial tissue. Anat Rec 1998;252:554–67. [62] Bolshakov A, Oliver WC, Pharr GM. Influences of stress on the measurement of mechanical properties using nanoindentation: Part II. Finite element simulations. J Mater Res 1996;11:760–8. [63] Hay JC, Bolshakov A, Pharr GM. Critical examination of the fundamental relations used in the analysis of nanoindentation data. J Mater Res 1999;14:2296–305. [64] Mencˇı´k J, Munz D, Quandt E, Weppelmann ER, Swain MV. Determination of elastic modulus of thin layers using nanoindentation. J Mater Res 1997;12:2475–84. [65] Shibata Y, He LH, Toda Y, Kataoka Y, Fujisawa N, Miyazaki T, et al. Micromechanical evaluation of mineralized multilayers. J Biomech 2008;41:3414–8. [66] Iwai-Yoshida M, Shibata Y, Wurihan, Suzuki D, Fujisawa N, Tanimoto Y, et al. Antioxidant and osteogenic properties of anodically oxidized titanium. J Mech Behav Biomed Mater 2012;13:230–6. [67] Jang IK, Tanaka R, Wurihan, Suzuki D, Shibata Y, Fujisawa N, et al. Nanomechanical properties and molecular structures of in vitro mineralized tissues on anodicallyoxidized titanium surfaces. Nanomedicine 2013;10: 629–37. [68] Shibata Y, Suzuki D, Wurihan, Yamada A, Maruyama N, Fujisawa N, et al. Lysyl oxidase like-2 reinforces unsatisfactory ossification induced by bone morphogenetic protein-2: relating nanomechanical properties and molecular changes. Nanomedicine 2013;9:1036–47. [69] Paschalis EP, Verdelis K, Doty SB, Boskey AL, Mendelsohn R, Yamauchi M. Spectroscopic characterization of collagen cross-links in bone. J Bone Miner Res 2001;16:1821–8. [70] Vora SR, Palamakumbura AH, Mitsi M, Guo Y, Pischon N, Nugent MA, et al. Lysyl oxidase propeptide inhibits

94

[71]

[72]

[73]

[74]

[75]

[76] [77]

[78]

[79]

[80]

[81]

[82]

[83]

[84]

[85]

[86]

[87]

[88]

[89]

journal of prosthodontic research 59 (2015) 84–95

FGF-2-induced signaling and proliferation of osteoblasts. J Biol Chem 2010;285:7384–93. Carden A, Morris MD. Application of vibrational spectroscopy to the study of mineralized tissues (Review). J Biomed Opt 2000;5:259–68. Dehring KA, Smukler AR, Roessler BJ, Morris MD. Correlating changes in collagen secondary structure with aging and defective type II collagen by Raman spectroscopy. Appl Spectrosc 2006;60:366–72. Dehring KA, Crane NJ, Smukler AR, McHugh JB, Roessler BJ, Morris MD. Identifying chemical changes in subchondral bone taken from murine knee joints using Raman spectroscopy. Appl Spectrosc 2006;60:1134–41. Xu C, Karan K, Yao X, Wang Y. Molecular structural analysis of noncarious cervical sclerotic dentin using Raman spectroscopy. J Raman Spectrosc 2009;40:1780–5. Bonifacio A, Beleites C, Vittur F, Marsich E, Semeraro S, Paoletti S, et al. Chemical imaging of articular cartilage sections with Raman mapping, employing uni- and multi-variate methods for data analysis. Analyst 2010;135:3193–204. Sikavitsas VI, Temenoff JS, Mikos AG. Biomaterials and bone mechanotransduction. Biomaterials 2001;22:2581–93. Burger EH, Klein-Nulend J. Mechanotransduction in bone – role of the lacuno-canalicular network. FASEB J 1999;13:S101–12. Klein-Nulend J, Van der Plas A, Semeins CM, Ajubi NE, Frangos JA, Nijweide PJ, et al. Sensitivity of osteocytes to biomechanical stress in vitro. FASEB J 1995;9:441–5. Knothe Tate ML, Knothe U, Niederer P. Experimental elucidation of mechanical load-induced fluid flow and its potential role in bone metabolism and functional adaptation. Am J Med Sci 1998;316:189–95. Weinbaum S, Cowin SC, Zeng Y. A model for the excitation of osteocytes by mechanical loading-induced bone fluid shear stresses. J Biomech 1994;27:339–60. Cook SD, Weinstein AM, Klawitter JJ. A three-dimensional finite element analysis of a porous rooted Co–Cr–Mo alloy dental implant. J Dent Res 1982;61:25–9. Knabe C, Howlett CR, Klar F, Zreiqat H. The effect of different titanium and hydroxyapatite-coated dental implant surfaces on phenotypic expression of human bone-derived cells. J Biomed Mater Res A 2004;71:98–107. Ausiello P, Franciosa P, Martorelli M, Watts DC. Effects of thread features in osseo-integrated titanium implants using a statistics-based finite element method. Dent Mater 2012;28:919–27. Baggi L, Cappelloni I, Di Girolamo M, Maceri F, Vairo G. The influence of implant diameter and length on stress distribution of osseointegrated implants related to crestal bone geometry: a three-dimensional finite element analysis. J Prosthet Dent 2008;100:422–31. Ding X, Zhu XH, Liao SH, Zhang XH, Chen H. Implant–bone interface stress distribution in immediately loaded implants of different diameters: a three-dimensional finite element analysis. J Prosthodont 2009;18:393–402. Ferraz CC, Anchieta RB, de Almeida EO, Freitas AC, Ferraz FC, Machado LS, et al. Influence of microthreads and platform switching on stress distribution in bone using angled abutments. J Prosthodont Res 2012;56: 256–63. Harirforoush R, Arzanpour S, Chehroudi B. The effects of implant angulation on the resonance frequency of a dental implant. Med Eng Phys 2014;36:1024–32. Kitagawa T, Tanimoto Y, Nemoto K, Aida M. Influence of cortical bone quality on stress distribution in bone around dental implant. Dent Mater J 2005;24:219–24. Pen˜arrocha-Diago MA, Flichy-Ferna´ndez AJ, AlonsoGonza´lez R, Pen˜arrocha-Oltra D, Balaguer-Martı´nez J,

[90]

[91]

[92]

[93]

[94]

[95]

[96]

[97]

[98] [99]

[100]

[101]

[102]

[103] [104]

[105]

[106]

Pen˜arrocha-Diago M. Influence of implant neck design and implant-abutment connection type on peri-implant health. Radiological study. Clin Oral Implants Res 2013;24:1192–200. Tanimoto Y, Hayakawa T, Nemoto K. Mode superposition transient dynamic analysis for dental implants with stress-absorbing elements: a finite element analysis. Dent Mater J 2006;25:480–6. Bahrami B, Shahrbaf S, Mirzakouchaki B, Ghalichi F, Ashtiani M, Martin N. Effect of surface treatment on stress distribution in immediately loaded dental implants – a 3D finite element analysis. Dent Mater 2014;30:e89–97. Huang HL, Hsu JT, Fuh LJ, Lin DJ, Chen MYC. Biomechanical simulation of various surface roughnesses and geometric designs on an immediately loaded dental implant. Comput Biol Med 2010;40:525–32. Jacobs JJ, Skipor AK, Patterson LM, Hallab NJ, Paprosky WG, Black J, et al. Metal release in patients who have had a primary total hip arthroplasty: a prospective, controlled, longitudinal study. J Bone Joint Surg Am 1998;80: 1447–58. Wen CE, Mabuchi M, Yamada Y, Shimojima K, Chino Y, Asahina T. Processing of biocompatible porous Ti and Mg. Scr Mater 2001;45:1147–53. Savadi RC, Agarwal J, Agarwal RS, Rangarajan V. Influence of implant surface topography and loading condition on stress distribution in bone around implants: a comparative 3D FEA. J Indian Prosthodont Soc 2011;11:221–31. Piotrowski B, Baptista AA, Patoor E, Bravetti P, Eberhardt A, Laheurte P. Interaction of bone–dental implant with new ultra low modulus alloy using a numerical approach. Mater Sci Eng C 2014;38:151–60. Al-Sayyed A, Deporter DA, Pilliar RM, Watson PA, Pharoah M, Berhane K, et al. Predictable crestal bone remodelling around two porous-coated titanium alloy dental implant designs. A radiographic study in dogs. Clin Oral Implants Res 1994;5:131–41. Cowin SC, Hegedus DH. Bone remodeling I: theory of adaptive elasticity. J Elast 1976;6:313–26. Huiskes R, Weinans H, Grootenboer HJ, Dalstra M, Fudala B, Slooff TJ. Adaptive bone-remodeling theory applied to prosthetic-design analysis. J Biomech 1987;20:1135–50. Lanyon LE. Functional strain in bone tissue as an objective and controlling stimulus for adaptive bone remodelling. J Biomech 1987;20:1083–93. O’Connor JA, Lanyon LE, MacFie H. The influence of strain rate on adaptive bone remodelling. J Biomech 1982;15:767–81. Huiskes R, Weinans H, Van Rietbergen B. The relationship between stress shielding and bone resorption around total hip stems and the effects of flexible materials. Clin Orthop 1992;12:4–34. Kopperdahl DL, Keaveny TM. Yield strain behavior of trabecular bone. J Biomech 1998;31:601–8. Van Rietbergen B, Huiskes R, Weinans H, Sumner DR, Turner TM, Galante JO. The mechanism of bone remodeling and resorption around press-fitted THA stems. J Biomech 1993;26:369–82. Boutroy S, Van Rietbergen B, Sornay-Rendu E, Munoz F, Bouxsein ML, Delmas PD. Finite element analysis based on in vivo HR-pQCT images of the distal radius is associated with wrist fracture in postmenopausal women. J Bone Miner Res 2008;23:392–9. Ito M, Nishida A, Koga A, Ikeda S, Shiraishi A, Uetani M, et al. Contribution of trabecular and cortical components to the mechanical properties of bone and their regulating parameters. Bone 2002;31:351–8.

journal of prosthodontic research 59 (2015) 84–95

[107] Newitt DC, Majumdar S, Van Rietbergen B, Von Ingersleben G, Harris ST, Genant HK, et al. In vivo assessment of architecture and micro-finite element analysis derived indices of mechanical properties of trabecular bone in the radius. Osteoporos Int 2002;13: 6–17.

95

[108] Eser A, Tonuk E, Akca K, Dard MM, Cehreli MC. Predicting bone remodeling around tissue- and bone-level dental implants used in reduced bone width. J Biomech 2013;46:2250–7. [109] Lin D, Li Q, Li W, Duckmanton N, Swain M. Mandibular bone remodeling induced by dental implant. J Biomech 2010;43:287–93.

A review of improved fixation methods for dental implants. Part II: biomechanical integrity at bone-implant interface.

The purpose of this article is to review the mechanical requirements of the tissue-implant interface and analyze related theories...
1MB Sizes 0 Downloads 9 Views