You are on page 1of 7

Journal of Biomechanics 45 (2012) 2241–2247

Contents lists available at SciVerse ScienceDirect

Journal of Biomechanics
journal homepage: www.elsevier.com/locate/jbiomech
www.JBiomech.com

A paradigm for the development and evaluation of novel implant topologies


for bone fixation: Implant design and fabrication
Heesuk Kang a, Jason P. Long b, Gary D. Urbiel Goldner a,b, Steven A. Goldstein a,b,c, Scott J. Hollister a,c,d,n
a
Department of Mechanical Engineering, the University of Michigan, USA
b
Department of Orthopaedic Surgery, the University of Michigan, USA
c
Department of Biomedical Engineering, the University of Michigan, USA
d
Department of Surgery, the University of Michigan, USA

a r t i c l e i n f o abstract

Article history: The future development of bio-integrated devices will improve the functionality of robotic prosthetic
Accepted 9 June 2012 limbs. A critical step in the advancement of bio-integrated prostheses will be establishing long-term,
secure fixation to the remnant bone. To overcome limitations associated with contemporary bone-
Keywords: anchored prosthetic limbs, we established a paradigm for developing and fabricating novel orthopedic
Orthopedic implant design implants undergoing specified loading. A topology optimization scheme was utilized to generate
Topology optimization optimal implant macrostructures that minimize deformations near the bone-implant interface.
Solid freeform fabrication Variations in implant characteristics and interfacial connectivity were investigated to examine how
Microstructure these variables influence the layout of the optimized implant. For enhanced tissue integration, the
optimally designed macroscopic geometry of a titanium (Ti)-alloy implant was further modified by
introducing optimized microstructures. The complex geometries of selected implants were successfully
fabricated using selective laser sintering (SLS) technology. Fabrication accuracy was assessed by
comparing volumes and cross-sectional areas of fabricated implants to CAD data. The error of fabricated
volume to CAD design volume was less than 8% and differences in cross sectional areas between SEM
images of fabricated implants and corresponding cross sections from CAD design were on average less
than 9%. We have demonstrated that this computational design method, combined with solid freeform
fabrication techniques, provides a versatile way to develop novel orthopedic implants.
& 2012 Published by Elsevier Ltd.

1. Introduction The concept of bone-anchored prosthetic limbs is not novel.


Currently, these devices are used by a limited number of patients
Contemporary robotic prostheses are capable of high-levels of with most having transfemoral amputations (Branemark et al.,
articulation and complex movements, but the usability of these 2001; Hagberg and Branemark, 2009). Complications with these
devices are limited by inadequate exchange of control and implant systems, including bone resorption around the anchor
sensory data between the user and prosthetic limb (Loeb, 2009). due to infection or due to stress-shielding of the surrounding
In the future, major advancements in functionality will likely bone (Tillander et al., 2010; Tomaszewski et al., 2010; Xu and
require permanent integration with the user’s body. A critical step Robinson, 2008), have yet to be adequately addressed. Addition-
in the development of a bio-integrated prosthesis will be estab- ally, current bone-anchored devices generally require a substan-
lishing technologies that enable long-term, secure fixation of the tial amount of remnant bone to structurally secure the prosthesis
devices to the bone in the limb remnant. This would allow for the (Branemark et al., 2001; Hagberg and Branemark, 2009), which
transfer of multi-axial and multi-directional loads generated limits these systems to a narrow range of anatomic sites and
during normal daily activity, and establish a structurally-secure clinical circumstances. These concerns highlight the need for
interface and anchor for the acquisition and transmission of further development of bone-secured prosthetic limbs and also
neural or muscular data. offer insight into the complex design requirements for an anchor
system that provides both robust mechanical function and
permits long-term preservation of the surrounding, structurally-
n
critical bone.
Corresponding author at: Department of Biomedical Engineering, University of
One approach to address these design complexities utilizes
Michigan, Lurie Biomedical Engineering Bldg. Rm 2208, 1101 Beal Ave. Ann Arbor,
MI 48109-2110, USA. Tel.: þ 1 734 647 9962; fax: þ1 734 647 0003. optimization strategies. For example, topology optimization in
E-mail address: scottho@umich.edu (S.J. Hollister). conjunction with the finite element (FE) method could be used to

0021-9290/$ - see front matter & 2012 Published by Elsevier Ltd.


http://dx.doi.org/10.1016/j.jbiomech.2012.06.011
2242 H. Kang et al. / Journal of Biomechanics 45 (2012) 2241–2247

optimize the layout of limited implant material in order to meet Table 1


specific performance criteria (Lin et al., 2004, 2007). This Linear elastic, isotropic material properties for finite element model.
approach could lead to the development of novel implant geo-
Component Young’s modulus (MPa) Poisson’s ratio
metries and topologies predicted to have robust mechanical
fixation under pre-determined loading conditions while limiting Trabecular bone 500 0.4
the amount of implant material. As part of this design strategy, it Fibrous tissue 1 0.45
is also critical that the implants can be feasibly manufactured and Titanium 110,000 0.33
75% Porous tantalum 3300 0.31
fabricated from material with high mechanical strength and
toughness to withstand repetitive loading over decades of use.
To address these issues, the aims of this study were: (1) to use created for a 451 wedge and then duplicated around the central vertical axis. The
topology optimization in conjunction with a FE model to develop mesh was generated using HyperMesh (HyperWorks 8.0, Altair Engineering Inc.,
implant designs for a favorable bone fixation response and (2) to Troy, MI) and consisted of 8-noded hexahedral and 6-noded pentahedral ele-
fabricate these optimized designs using available technologies. ments. All elements were assigned linear elastic, isotropic material properties
(Table 1).
To develop a range of implant designs, we investigated variations in implant
characteristics and interfacial connectivity. Changes to implant characteristics
2. Materials and methods included different metals, Ti-alloy or 70% porous tantalum (Table 1), that are
commonly used in contemporary orthopedic implants, and variations in design
Topology optimization is an approach to optimize the material layout within a domain size (implant dimensions were constant, but the design domain made up
specified design domain, such that the resulting structure meets specific perfor- 67%, 42% or 25% of the overall radial dimension). Interfacial connectivity was
mance criteria (Bendsøe and Kikuchi, 1988). In general, this optimization method varied to simulate potential differences in bone-implant integration that may
is executed in conjunction with FE methods, on which a density value is assigned develop during the early post-operative period. Three interface variations were
to each element that represents the structural topology. The density value ranges examined, including fibrous tissue or trabecular bone material properties assigned
from 0 to 1, with 0 indicating no material (void) and 1 indicating a solid. to the entire interface zone. For the third variation, we assigned fibrous tissue to
Introduction of a material interpolation law such as the homogenization method the circumferential implant interface zone and trabecular bone to the bottom
or solid isotropic material penalization (SIMP) relates the density values to the implant interface zone. For all models, implant-tissue surfaces were displacement
element material properties (Sigmund, 1994). In this way, the intermediate values compatible. In total, we analyzed 18 different FE models.
can indicate partial material that requires post-processing to establish a clear The optimized implant structure will be highly dependent upon its loading
solid-void boundary. The optimization algorithm evolves the material density environment. A bone anchor for a bio-integrated prosthetic limb will undergo
distribution such that the objective of the optimization problem is minimized complex loading during daily use, and the loading environment will depend upon
under specified boundary conditions and volume constraints. a number of factors, including anatomic site and patient activity level. As an initial
The objective function in this particular optimization scheme was to minimize iteration for establishing our design paradigm, loading was simplified to uni-axial
the compliance of the bone-implant system, which was achieved by altering the forces (100 N tension/compression) applied to the top of the loading post. The
distribution of implant material. In doing so, deformations near the bone-implant model was fully constrained by assigning a zero displacement boundary condition
interface were minimized and the stiffness at this interface region was maximized. to the outer surfaces (i.e. sides and base) of the bone.
As an initial investigation towards minimizing the amount of implant material, we The optimization problems were solved using OptiStruct (HyperWorks 8.0,
constrained the volume of the optimized macrostructure to compose no more Altair Engineering Inc., Troy, MI), a commercial topology optimization solver.
than 50% of the design domain volume. Following optimization convergence, a threshold value was applied to the design
Our FE model had a cylindrical implant design domain surrounded by a bed of domain’s density distribution to define the structural boundary of each implant.
homogeneous trabecular bone (Fig. 1), which represents a generic metaphyseal Regions with density values higher than the threshold were considered as solid;
region rather than a specific anatomic site. With this approach, we sought to regions below the threshold were considered void. Then, the material volume
establish a design paradigm rather than create an implant for clinical application. fraction or porosity of the designed implant was evaluated to determine if it met
Nevertheless, for potential translation, our model could represent a single the volume fraction constraints of the optimization scheme.
structural projection that would work in concert with similar projections to meet Following the solution of the various optimization problems, implant designs
the mechanical requirements of an implant system (Goldstein et al., 1991; were selected for fabrication and future in vivo evaluation. For our initial selection,
Matthews and Goldstein, 1992). This concept would be particularly attractive in we chose an optimized design made of Ti-alloy. As a subsequent design process,
cases where a patient has a limited length of remnant bone for attachment of a this selected design was modified to have a porous architecture by replacing
prosthetic anchor. material layout with appropriate microstructures to further enhance the integra-
Other aspects of our FE model included an interfacial zone between the tion of bone and implant. Specifically, the density distribution was segmented into
implant and bone with a loading post attached to the top of the implant. The high and low density regions, which served as a map to place specific micro-
design domain was defined as a hollow cylinder to maintain implant material near structures (Hollister and Lin, 2007; Lin et al., 2004). The low and high porosity
the loading post. To represent the entire cylindrical model, an FE mesh was microstructures used for this mapping were designed to be optimally stiff and
permeable, as per isotropic cross-property bounds (Gibiansky and Torquato,
1996). To do this, a multi-objective topology optimization code was used, and
the microstructures were optimized based upon bulk modulus and diffusivity
properties (Kang et al., 2010). The low and high density microstructures had
porosities of 63% and 33%, respectively (Fig. 2). The overall porosity of the scaffold
was constrained to be 50%.
For fabrication, optimized implant dimensions were set to 6.4 mm diameter
by 6.4 mm height. The shape of the loading post was modified, such that a narrow
post diameter was maintained at the junction with the optimized structure but
flared away from the implant to a diameter of 6.4 mm. The additional post
material allowed for the future placement of a drilled and tapped hole for in vivo
evaluation. Overall implant height was set to 12.7 mm (including the loading
post). Implants were fabricated from extra low interstitial (ELI)-grade Ti–6Al–4V
alloy by Synergeering Group (Farmington Hills, MI) using a selective laser
sintering (SLS) system (EOSINT M 270, EOS of North America Inc., Novi, MI).
Implants were not annealed, which could refine structure and alter material
properties (Hollander et al., 2006).
The manufacturing accuracy of the solid freeform fabricated implants was
assessed by two methods. First, micro-computed tomography (microCT) was used
to measure the volume fraction of implant material with respect to the cylindrical
design domain. To do this, implants were scanned with an eXplore Locus SP
microCT system (GE Healthcare Preclinical Imaging, London, Ontario) using
Fig. 1. The finite element model of the bone-implant system for topology 80 kVp X-ray potential, 6000 ms exposure time and a 29 mm isotropic voxel size.
optimization was cylindrical and consisted of a hollow design domain with a Metallic artifacts were minimized by aligning the axisymmetric implants with the
central loading post surrounded by trabecular bone. An interfacial tissue zone was specimen holder’s rotation axis. To measure volume fraction, MicroView (v 2.2, GE
also included between design domain and surrounding trabecular bone. Healthcare Preclinical Imaging) was used to generate a cylindrical region of
H. Kang et al. / Journal of Biomechanics 45 (2012) 2241–2247 2243

Fig. 3. The topology optimization distributed more material in high stress regions
such that the stiffness of the system was maximized for a given material volume
fraction. In this example, the entire interfacial zone was assumed to be
trabecular bone.

Fig. 2. Optimized microstructures were used for generating a hierarchical porous


implant by mapping the density distribution from the macrostructure topology
optimization. The effective mechanical and mass transport properties of the
microstructures are close to theoretical upper bounds, indicating that these
microstructures simultaneously provide maximized mechanical and mass trans-
port properties for a given porosity. In this figure, K refers to the bulk modulus of
the optimized microstructure and K1 refers to the bulk modulus of a solid
microstructure (i.e. zero porosity). Similarly, D refers to the diffusivity of the
optimized microstructure and D2 refers to the diffusivity of void space (i.e. 100%
porosity).

interest (ROI) that matched the cylindrical design domain of the optimization
scheme (6.4 mm diameter by 6.4 mm height). Implant material was thresholded
to calculate the volume fraction of the ROI. For the second method of assessing
manufacturing accuracy, cross-sections from porous optimized implants were
imaged with a backscatter scanning electron microscope (SEM). To do this, two Fig. 4. Implant material, design domain size and interfacial connectivity were
implants were embedded in methylmethacrylate and cut with a diamond saw varied to explore the effects of these variables upon the topology optimization
(IsoMet Low Speed Saw, Buehler LTD, Lake Bluff, IL) generating roughly 1000 mm results. Each row of implants represents a different design domain size (67%, 42%
thick sections; one implant was sectioned orthogonal to the long axis (transverse or 25% of the overall radial dimension). (A) and (D) are with fibrous tissue along
sections) and the other parallel (longitudinal sections). Sections were wet polished the entire interface, (B) and (E) are with fibrous tissue interface along the implant
to a final thickness of 800–900 mm, carbon-coated and imaged at 10  magnifica- sides and a bony interface along the base, and (C) and (F) are with bony tissue
tion with an Amray SEM (model 1810, KLA-Tencor Corp., Milpitas, CA). Using along the entire interface.
Matlab (v 7.11, The MathWorks Inc., Natick, MA), digital SEM cross-sections were
then quantitatively compared to matched CAD cross-sections by overlaying
images at a pixel size of 18 mm. Each implant pixel was then categorized as
overlap (both SEM and CAD), SEM only or CAD only. For each cross-section, two introducing a fibrous tissue zone around the design domain, the
error percentages were calculated: (1) excess material was indicated by the pixel hollowed space filled with more implant material to increase the
number of SEM only and (2) deficient material by the pixel number of CAD only. overall stiffness of the bone-implant system. Varying implant
Both error percentages were normalized by the number of overall CAD pixels material only affected the topology when there was direct bone
within the cross-section.
apposition along the entire interface (i.e. both sides and bottom).
Specifically, the titanium implant had a large hollow base without
any implant material, while the porous tantalum implant had a
3. Results central column within this hollow region that connected the
implant and bone along the base of the design domain. Variations
To minimize total deformation under uni-axial loading, the in the size of the design domain had limited effects on topology.
optimization algorithm assigned higher density values to angled For fabrication, a candidate design was chosen that repre-
elements that generally formed a bell-shaped topology (Fig. 3). sented a titanium implant with a large design domain and
These higher density regions corresponded to finite elements trabecular bone interface (no fibrous tissue). As an additional
under high von Mises stresses. The stress distribution revealed design step, this solid macrostructure was transformed into a
that load could be transmitted to the bone along the side and hierarchical scaffold by mapping microstructure sub-units to the
bottom of the design domain. design domain’s density distribution (Fig. 5). This added process
Applying the FE topology optimization to the permutations of created a porous implant structure designed to facilitate bone
implant characteristics and interface connectivity resulted in the infiltration and improve implant fixation. The solid, macrostruc-
development of 18 macrostructural implant designs (Fig. 4). In ture and porous scaffold designs were then converted to CAD
general, many of the model permutations resulted in subtle format for fabrication (Fig. 6).
changes to the implant’s topology. The most substantial effect Both optimized implant designs were successfully fabricated
on design topology was associated with a fibrous or bone tissue using SLS technology (Fig. 7). In general, the topology of the
interface. Without a fibrous tissue zone (i.e. direct bone integra- fabricated implants matched the CAD data; however, some of the
tion), the topology optimization tended to develop a bell-shaped smaller pores associated with the porous design were filled with
implant design with a hollow region near the base. Upon excess Ti-alloy. Volume fractions of the manufactured implants,
2244 H. Kang et al. / Journal of Biomechanics 45 (2012) 2241–2247

4. Discussion

To improve functionality of prosthetic limbs, there are efforts


to develop bio-integrated devices. An important step in advancing
this technology will be establishing long-term, secure fixation of
the prosthesis to remnant bone. To overcome limitations asso-
ciated with contemporary bone-anchor systems, we established a
paradigm for designing, fabricating and evaluating novel ortho-
pedic implant topologies that will maintain long-term, robust
fixation in bone while undergoing loading. In the present study,
we used topology optimization to explore potential implant
geometries that minimize deformation at the bone-implant inter-
face with limited implant material. The versatility of the FE
topology optimization design scheme was demonstrated by
examining the effects of variations in implant material, design
domain size and interfacial connectivity upon the optimized
results. As an additional design step, optimal microstructures
were utilized to replace the global material density distribution
determined by the topology optimization. This created a macro–
micro implant structure that was designed to encourage deeper
bone infiltration and establish better implant-bone integration.
Selected implants were then fabricated by SLS of medical-grade
Ti-alloy with overall dimensions of 6.4  12.7 mm. Given the
small implant size, this fabrication technology proved to be
Fig. 5. Integration of the optimized macroscopic topology with the optimized accurate with errors assessed by volume and cross-sectional
microstructural unit cells created a hierarchical porous implant. Microstructures shape generally within710%.
with 63% and 33% porosities replaced low and high density regions, respectively. An advantage of topology optimization is that it can be used
directly in conjunction with bone mechano-regulatory models for
developing orthopedic implants. In the present study, the opti-
mization objective was to minimize the total compliance of the
bone-implant system, which was intended to minimize tissue
strains near the bone-implant interface. This objective function
was motivated by previous research that demonstrated bone-
implant integration is related to the local mechanical environ-
ment, which in turn can be affected by implant surface geometry
(Simmons et al., 2001a, 2001b). For example, Simmons et al.
(2001a) suggested 3% distortional strain as a threshold level,
above which local bone formation is inhibited, leading to poor
integration within an early healing period. Optimal implant
geometry could be designed by limiting the strain magnitudes
at the bone-implant interface. Although we did not assign con-
straints directly on the strain levels, all of our optimal implant
designs showed less than 1% strain when minimizing compliance.
Thus, we expect early interfacial bone formation during in vivo
evaluation of our optimally designed implants.
The integration of macro–micro topology optimized structures
can bring additional versatility in the design of implant struc-
tures. Microstructural topology optimization has been utilized to
Fig. 6. Designs selected for fabrication were a solid design derived directly from find optimal microstructures by defining the design domain
the macroscopic topology optimization and a hierarchical scaffold design gener- within the representative microstructure space (Sigmund, 1994).
ated from optimal microstructures.
The macroscopic or average properties are then evaluated
using the homogenization method by solving characteristic
equations within the microstructural domain (Gründemann,
as measured from microCT, were similar to CAD volume fractions. 1987). At each optimization iteration, material density at the
More specifically, the manufactured solid design had a lower microstructure is updated to minimize the error between target
volume fraction than the CAD image (34.4% versus 37.4%), while properties and current properties at the macroscopic level.
volume fraction of the manufactured porous design was higher Optimal microstructures can then be designed such that the
than the CAD image (51.7% versus 50.3%). Seven cross-sections equivalent macroscopic properties match those given as design
(five transverse and two longitudinal) of the porous implant were targets. Furthermore, multiple properties, such as mechanical
imaged with SEM, which showed that the hierarchical scaffold and mass transport, can be considered simultaneously in
was successfully fabricated (Fig. 8). However, dimensional devia- the microstructure design (Challis et al., 2008; Kang et al.,
tions, due to excess and absent material, were identified at all 2010). In addition, several microstructures can be tailored to
cross-sections (Table 2). Error percentages for excess material have different properties at the same porosity (Kang et al., 2010).
ranged from 5.5% to 11.8%, and absent material ranged from 3.3% This allows us to design porous implants with varying mechanical
to 12.8%. Of note, many internal pore structures had excess and mass transport properties for a given macroscopic material
material resulting in smaller pores than designed. distribution.
H. Kang et al. / Journal of Biomechanics 45 (2012) 2241–2247 2245

Fig. 7. Scanning electron microscopy (SEM) images show implants fabricated with selective laser sintered (SLS) technology.

Fig. 8. For assessing fabrication accuracy, implant cross-sections were imaged with SEM and then overlayed with CAD cross-sections.
2246 H. Kang et al. / Journal of Biomechanics 45 (2012) 2241–2247

Table 2
Comparison of implant cross-sections (Fabricated versus CAD).

Section Average

1 (Long.) 2 (Long.) 3 (Trans.) 4 (Trans.) 5 (Trans.) 6 (Trans.) 7 (Trans.)

% Excess Material (SEM only) 11.7 7.0 8.8 11.4 5.3 8.7 9.5 8.9
% Deficient Material (CAD only) 4.5 6.2 4.8 2.6 12.6 10.7 6.6 6.9

During the optimization process, implant material was removed in afterwards by using additional design and analysis tools. There
low von Mises stress regions and assumed to be replaced with void. are also limitations associated with the SLS technology we used to
This would be equivalent to the initial post-operative period where fabricate two optimized implant designs. In general, the small
bone ingrowth would have not yet occurred. Alternatively, one pores of the hierarchical scaffold were smaller than designed due
could modify the material interpolation such that lower modulus to excess implant material. These pores were designed to be
(trabecular bone) and higher modulus (implant metal) materials are 350 mm, which is near the fabrication resolution of contemporary
incorporated into the design domain. This would assume full integra- SLS systems. Additionally, material properties of the implant will
tion of the host bone to the implant surface. However, since the be dependent upon SLS fabrication parameters (e.g. powder size
modulus of bone is less than that of the implant materials, the effect and laser temperature). Post-processing, like annealing, can be
of introducing low stiffness material into the design domain may not used to alter material properties, such as toughness and fatigue
significantly impact the optimization results. life (Hollander et al., 2006).
There are several limitations associated with the implant
design scheme, topology optimization and SLS fabrication tech-
nology. Our FE model was a generic representation of a small 5. Conclusions
bone-implant system exposed to simplified boundary conditions.
Only uni-axial loads were applied to the implant for this initial We successfully demonstrated the development of implant
iteration of our model. While this is not a complex loading designs for favorable bone fixation using topology optimization.
condition, our goal at this stage was to establish a design This approach allows for designing novel orthopedic implants for
paradigm for generating and evaluating novel implant structures; specific loading environments that have intricate porous archi-
we did not seek to create an implant for clinical use. Focusing on tectures for tissue integration. Candidate implants were then
uni-axial loads also enabled us to evaluate our fabricated successfully fabricated using SLS technology with medical-grade
implants in a companion experiment that incorporated in vivo Ti-alloy. The fabricated implants closely matched CAD data,
application of cyclic, tension-compression loads (Long et al., generally within 7 10%. The next step will be testing our implants
under review). Another limitation of our FE model was using in an in vivo environment to identify key aspects of the designs
the 451 wedge to represent the entire 3D structure. This was done that promote osseointegration.
to improve the computational efficiency of the optimization
algorithm; however, the wedge mesh (and the associated bound- Conflicts of interest
ary conditions) created stress artifacts near the edges. In some
cases, these artifacts affected the topology of the optimized The authors have no conflicts of interest to disclose.
implant. For example, the small radial projections near the top
of the solid design fabricated from Ti-alloy are artifacts
(Figs. 6 and 7). In all cases, however, these artifacts were small Acknowledgments
relative to the global optimized structure. While there were
minor issues with our design scheme, it is important to note that This work was supported by a Multi-disciplinary University
the FE model could be modified and refined to develop other Research Initiative (MURI) from the Army Research Office (Pro-
implant structures, in which case the framework of our design posal 50376-LS-MUR, Grant W911NF-06-0218) as well as the
paradigm is still applicable. Regarding limitations of topology National Institutes of Health (Grants 5T90-DK070071, 5RO1-
optimization methods, future iterations of our optimization AR051504 and 1R01AR053379).
scheme could benefit from next generation software that would
permit modeling of more complex physical phenomena. For References
example, topology optimization typically requires linear elasti-
city; nonlinear material properties, nonlinear interface conditions Bendsøe, M.P., Kikuchi, N., 1988. Generating optimal topologies in structural
and large deformations may not be correctly modeled. In this design using a homogenization method. Computer Methods in Applied
study, we applied relatively small magnitude loads, which led to Mechanics 71, 197–224.
Branemark, R., Branemark, P.I., Rydevik, B., Myers, R.R., 2001. Osseointegration in
small displacements. In this situation bone is generally consid- skeletal reconstruction and rehabilitation: a review. Journal of Rehabilitation
ered as linear elastic material (Keaveny et al., 1994). Another Research and Development 38, 175–181.
limitation of current topology optimization algorithms is the Challis, V.J., Roberts, A.P., Wilkins, A.H., 2008. Design of three dimensional
isotropic microstructures for maximized stiffness and conductivity. Inter-
inability to account for time-dependent history of the mechanical national Journal of Solids and Structures 45, 4130–4146.
environment. In general, mechano-regulatory models relate Gibiansky, L.V., Torquato, S., 1996. Connection between the conductivity and bulk
mechanical loading history to bone maintenance, so developing modulus of isotropic composite materials. Proceedings of the Royal Socety of
London A: Mathematical, Physical and Engineering Sciences 452, 253–283.
algorithms that account for time-dependent mechanical stimuli
Goldstein, S.A., Matthews, L.S., Kuhn, J.L., Hollister, S.J., 1991. Trabecular bone
may improve the utility for orthopedic implant design. Overall, it remodeling: an experimental model. Journal of Biomechanics 24 (1), 135–150.
is important to note that topology optimization is primarily useful Gründemann, H., 1987. Homogenization techniques for composite media. In:
for developing structures during the early stages of the design Sanchez-Palencia, E., Zaoui, A., (Eds.), Homogenization techniques for compo-
site media, Lectures Delivered at the Proceedings of the International Center
process. In other words, results from topology optimization for Mechanical Sciences, CISM, Udine, Italy, 1–5 July, 1985. Springer-Verlag,
should serve as initial designs, and fine tuning can be executed Berlin.
H. Kang et al. / Journal of Biomechanics 45 (2012) 2241–2247 2247

Hagberg, K., Branemark, R., 2009. One hundred patients treated with osseointe- Long, J.P., Hollister, S.J., Goldstein, S.A., A paradigm for the development and
grated transfemoral amputation prostheses—rehabilitation perspective. Jour- evaluation of novel implant topologies for bone fixation: in vivo evaluation.
nal of Rehabilitation Research and Development 46, 331–344. Journal of Biomechanics, under review.
Hollander, D.A., von Walter, M., Wirtz, T., Sellei, R., Schmidt-Rohlfing, B., Paar, O., Matthews, L.S., Goldstein, S.A., 1992. The prosthesis-bone interface in total knee
Erli, H.J., 2006. Structural, mechanical and in vitro characterization of indivi- arthroplasty. Clinical Orthopaedics and Related Research, 50–55.
dually structured Ti–6Al–4V produced by direct laser forming. Biomaterials Sigmund, O., 1994. Materials with prescribed constitutive parameters—An inverse
27, 955–963. homogenization problem. International Journal of Solids and Structures 31,
Hollister, S.J., Lin, C.Y., 2007. Computational design of tissue engineering scaffolds. 2313–2329.
Computer Methods in Applied Mechanics 196, 2991–2998. Simmons, C.A., Meguid, S.A., Pilliar, R.M., 2001a. Mechanical regulation of localized
Kang, H., Lin, C.Y., Hollister, S.J., 2010. Topology optimization of three dimensional and appositional bone formation around bone-interfacing implants. Journal of
tissue engineering scaffold architectures for prescribed bulk modulus and Biomedical Materials Research 55, 63–71.
diffusivity. Structural and Multidisciplinary Optimization 42, 633–644. Simmons, C.A., Meguid, S.A., Pilliar, R.M., 2001b. Differences in osseointegration
Keaveny, T.M., Guo, X.E., Wachtel, E.F., McMahon, T.A., Hayes, W.C., 1994. rate due to implant surface geometry can be explained by local tissue strains.
Trabecular bone exhibits fully linear elastic behavior and yields at low strains. Journal of Orthopedic Research 19, 187–194.
Journal of Biomechanics 27, 1127–1136. Tillander, J., Hagberg, K., Hagberg, L., Branemark, R., 2010. Osseointegrated
Lin, C.Y., Hsiao, C.C., Chen, P.Q., Hollister, S.J., 2004. Interbody fusion cage design titanium implants for limb prostheses attachments: infectious complications.
using integrated global layout and local microstructure topology optimization. Clinical Orthopaedics and Related Research 468, 2781–2788.
Spine 29, 1747–1754. Tomaszewski, P.K., Verdonschot, N., Bulstra, S.K., Verkerke, G.J., 2010. A compara-
Lin, C.Y., Wirtz, T., LaMarca, F., Hollister, S.J., 2007. Structural and mechanical tive finite-element analysis of bone failure and load transfer of osseointe-
evaluations of a topology optimized titanium interbody fusion cage fabricated by grated prostheses fixations. Annals of Biomedical Engineering 38, 2418–2427.
selective laser melting process. Journal of Biomedical Materials Research A 83, Xu, W., Robinson, K., 2008. X-ray image review of the bone remodeling around an
272–279. osseointegrated trans-femoral implant and a finite element simulation case
Loeb, G.E., 2009. Taking control of prosthetic arms. JAMA 301, 670–671. study. Annals of Biomedical Engineering 36, 435–443.

You might also like