Sie sind auf Seite 1von 7

In Gwun Jang

e-mail: jangin@me.queensu.ca

Il Yong Kim1
e-mail: iykim@me.queensu.ca
Department of Mechanical and Materials
Engineering,
Queens University,
McLaughlin Hall 221,
130 Stuart Street
Kingston, ON, K7L 3N6, Canada

Byung Man Kwak


Department of Mechanical Engineering,
Korea Advanced Institute of Science and
Technology,
373-1 Guseong-dong, Yuseong-gu
Daejeon 305-701, South Korea
e-mail: bmkwak@khp.kaist.ac.kr

Analogy of Strain Energy Density


Based Bone-Remodeling
Algorithm and Structural
Topology Optimization
In bone-remodeling studies, it is believed that the morphology of bone is affected by its
internal mechanical loads. From the 1970s, high computing power enabled quantitative
studies in the simulation of bone remodeling or bone adaptation. Among them, Huiskes et
al. (1987, Adaptive Bone Remodeling Theory Applied to Prosthetic Design Analysis, J.
Biomech. Eng., 20, pp. 11351150) proposed a strain energy density based approach to
bone remodeling and used the apparent density for the characterization of internal bone
morphology. The fundamental idea was that bone density would increase when strain (or
strain energy density) is higher than a certain value and bone resorption would occur
when the strain (or strain energy density) quantities are lower than the threshold. Several
advanced algorithms were developed based on these studies in an attempt to more accurately simulate physiological bone-remodeling processes. As another approach, topology
optimization originally devised in structural optimization has been also used in the computational simulation of the bone-remodeling process. The topology optimization method
systematically and iteratively distributes material in a design domain, determining an
optimal structure that minimizes an objective function. In this paper, we compared two
seemingly different approaches in different fieldsthe strain energy density based boneremodeling algorithm (biomechanical approach) and the compliance based structural
topology optimization method (mechanical approach)in terms of mathematical formulations, numerical difficulties, and behavior of their numerical solutions. Two numerical
case studies were conducted to demonstrate their similarity and difference, and then the
solution convergences were discussed quantitatively. DOI: 10.1115/1.3005202
Keywords: bone remodeling, bone adaptation, topology optimization, computational
simulation

Introduction

In the field of bone remodeling, it is believed that bone has


self-optimizing capabilities and therefore is able to control its
mass and structure according to its mechanical demands. From the
1970s, the advance of computer capabilities and numerical stress
analysis methods has allowed for the quantitative study of bone
remodeling. Cowin and Hegedus 1 proposed the first quantitative bone-remodeling equations based on continuum mechanics.
Fyhrie and Carter 2 postulated that bone adaptively changes its
structure and density in response to its stress and strain state.
Huiskes et al. 3 proposed another remodeling algorithm based
on the assumption that strain energy density SED acts as the
stimulus for bone density change. In these approaches, the rate of
the bone density change is determined on the basis of the difference between the remodeling stimulus e.g., stress, strain, and
strain energy density and its target value. A number of different
approaches have been proposed afterwards 47.
In the aforementioned studies, the behavior of the bone structure is described by a set of differential equations, and the bone
may be considered as a nonlinear system with a certain number of
degrees of freedom, which is the same as the number of finite
elements in the analysis domain. It is interesting to note that these
1
Corresponding author.
Contributed by the Bioengineering Division of ASME for publication in the JOURNAL OF BIOMECHANICAL ENGINEERING. Manuscript received January 14, 2008; final
manuscript received July 18, 2008; published online November 26, 2008. Review
conducted by Ellen M. Arruda. Paper presented at the 21st Canadian Congress of
Applied Mechanics, 2007.

Journal of Biomechanical Engineering

elementwise differential equations may be replaced by a domainwide objective function in optimization. A similar relationship is
found in Newtonian mechanics, in which the equilibrium state of
objects is obtained by solving the simultaneous governing equations, and the same equilibrium can also be determined by minimizing the energy state of the objects 8.
In mechanical engineering, there has been a significant progress
in structural optimization, which is often classified into size,
shape, and topology optimization. Among these methods, topology optimization iteratively redistributes material in the domain to
determine an optimal material arrangement 911. Since the
1990s, topology optimization has been successfully applied to the
simulation of bone remodeling. Hollister et al. 12 used the homogenization method to obtain microtrabecular bone architecture.
Fernandes and Rodrigues 13 used a weighted sum of strain energy and mass as the objective function in topology optimization
to simulate bone remodeling around cementless hip stems. Bagge
14 used topology optimization with compliance minimization to
determine the initial material distribution for proximal femur.
Huiskes 15 summarized this observation: It has been shown
that computer algorithms for material design optimization, when
given the opportunity, show the tendency to produce trabecular
patterns, similar to bone. These studies provided insight into
bone as an optimized structure, but they rested on the assumption
without any solid foundation of theoretical proof.
Even though it is well known by experience that the two seemingly different approaches in different fieldsthe elementwise
bone-remodeling algorithms and the structural topology optimization with a domain-wide objective functionproduce similar re-

Copyright 2009 by ASME

JANUARY 2009, Vol. 131 / 011012-1

Downloaded From: http://biomechanical.asmedigitalcollection.asme.org/ on 03/05/2015 Terms of Use: http://asme.org/terms

sults, it has not been thoroughly studied how these two approaches are compared explicitly: In terms of mathematical
formulations, numerical difficulties and their remedies, and behavior of their numerical solutions. Only a few studies tried to
derive the relationships between the two methods. Harrigan and
Hamilton 16 formulated a global functional for the simulation of
bone remodeling as the weighted sum of mass and the total strain
energy. They showed that the role of the functional was equivalent
to that of the remodeling equation developed by Huiskes and coworkers 5,6. Subbaraya and Bartel 17 also suggested the same
global functional for local bone-remodeling phenomenon. In the
aforementioned studies, the proposed global functional was totally
equivalent to the bone remodeling equation by Huiskes et al.
However, the simulations based on this functional are unconstrained optimization tasks, and they are different from the typical
topology optimization with a mass constraint that has been used
most widely in various applications including bone-remodeling
simulation. The analogy between this standard topology optimization with a mass constraint and the stain energy based boneremodeling equations has not been studied, and the relationships
between the topology optimization parameters and the boneremodeling equation variables are not yet identified.
In this paper, we formulated the typical form of topology optimization starting from the global functional by Harrigan and
Hamilton 16 and investigated the similarities and differences
between this topology optimization and the strain energy density
based approach in terms of mathematical formulations and numerical implementations. We then determined the matching relationships among the parameters used in the topology optimization
and bone-remodeling algorithm. Two numerical examples were
solved using the two approaches and the behavior of each numerical convergence was discussed.

2 Similarity of Solution Methods of Bone Remodeling


and Topology Optimization
In the 1990s, macroscopic continuum models for bone remodeling were developed, in which the entire domain was discretized
into a number of design cells, and mechanical variables and biological responses were spatially averaged to yield appropriate apparent variables over the region of each design cell. Indeed all the
bone-remodeling algorithms aforementioned were developed
based on this approach. It is interesting to note that the same
approach has been used in topology optimization. This cell-based
approach has been very widely used in topology optimization.
The second similarity between bone remodeling and topology
optimization is the relationship between the density and Youngs
modulus E. In topology optimization with the solid isotropic material with penalization SIMP approach 10, which is one of the
most popular methods nowadays, an artificial relationship between the density and Youngs modulus is used based on a power
law. The relationship between and E of bone in bone remodeling
5,6 is the same as that of the SIMP model in topology optimization; the mathematical expression is
E = c

where E is Youngs modulus, is the density, and c and are the


constants to be selected.
Third, we gain insight into the behavior of a major state variable over computational iterations in bone-remodeling simulations
and topology optimization. Energy stored in each cell is used as a
state variable in the SED-based bone remodeling equations. As
remodeling progresses, the distribution of the stored energy over
the entire domain becomes more uniform. In structural topology
optimization, the ideal optimum design has the fully stressed
state where all the parts of a structure can be used most efficiently.
In an actual case, however, the optimum solution would not have
a completely constant stress distribution, but the variance in stress
distribution always decreases as optimization progresses.

Finally, it is also worth noting that the SED-based boneremodeling approaches and the topology optimization methods
have the same numerical difficultiescheckerboard patterns and
solution dependency on mesh resolutionand moreover the remedies used in each area are similar although the numerical techniques were developed independently. Weinans et al. 5 found
checkerboard patterns and mesh dependency in their numerical
bone-remodeling simulations. They used the terms patchwork effect for checkerboard pattern and chaotic phenomena for mesh
dependency. To overcome checkerboard patterns in bone remodeling, Jacobs et al. 18 suggested the use of quadratic elements
over bilinear elements and also the use of node-based approach;
these two remedies were also used in the same way in topology
optimization 19,20. More active schemes were developed to resolve the numerical difficulties: Mullender et al. 6 introduced the
influence area of sensors, and similar techniques were developed in topology optimization with the name of sensitivity filtering 21. Mathematical equations of these techniques are presented in Sec. 3.

Mathematical Formulations

3.1 Bone-Remodeling Simulation Using the Strain Energy


Density Based Approach. It is believed that the mass and form of
bone are determined based on the two processesosteoclast resorption and osteoblast formationand the processes are modulated by external loads through osteocytic sensing and signaling.
To consider this biological phenomenon in numerical simulation,
Weinans et al. 5 assumed that a difference between the actual
SED, S = Sx , y , z, and a reference SED, k = kx , y , z, at the same
location was the driving force for adaptation of the apparent density. The stimulus was then approximated by U / , where U is the
apparent SED in the bone and is the apparent density. They
proposed a bone-remodeling equation for multiple load cases:

d j
Ua,j
k ,
=B
dt
j

0 j upper

where B and k are constants, and Ua,j represents the apparent


strain energy density of the jth element. Here,
n

Ua,j =

1
Ui,j
n i=1

where Ui,j is the apparent strain energy density of the jth element
for load case i, and n is the number of load cases. The relation
between Youngs modulus and the density was taken as in Eq. 1.
Usually, in Eq. 1 has a value of 2 or 3. The remodeling process
is considered to converge when d / dt becomes zero in Eq. 2 or
when the density reaches the lower or upper bound.
To make the bone-remodeling process more physiological i.e.,
no checkerboards or no mesh dependency, Mullender et al. 6
introduced a spatial influence function fx to Eq. 2:
d
=
dt

f x
N

Ua,j

j=1

k ,

0 upper

where N is the number of sensor cells, and f jx = edistx,j/D with


dist x , j the distance between sensor j and location x. They assumed that sensor cells were uniformly distributed over its volume.
3.2 Structural Topology Optimization With Compliance
Minimization. In this section, we will show that the SED-based
bone-remodeling algorithm, which was devised in biomechanics,
can be matched mathematically with the topology optimization
formulation. We will then reveal the relations among the constants
used in the SED-based bone-remodeling algorithm and topology
optimization. We will also discuss how the sensitivity filtering in
topology optimization is compared to the spatial influence func-

011012-2 / Vol. 131, JANUARY 2009

Downloaded From: http://biomechanical.asmedigitalcollection.asme.org/ on 03/05/2015 Terms of Use: http://asme.org/terms

Transactions of the ASME

tion in the SED-based bone-remodeling algorithm.


Harrigan and Hamilton 16 formulated a global functional for
the simulation of bone remodeling as the weighted sum of mass
and the total strain energy. They showed that the role of the functional was equivalent to that of the remodeling equation developed by Huiskes and co-workers 5,6. Now, starting from this
global functional as an objective function, we formulate an unconstrained optimization problem under multiple loads as

minimize =

where ui is the displacement vector for load case i, n is the number of load cases, K is the global stiffness matrix of the structure,
j is the jth element density, j is the jth element volume, and N
is the total number of elements. We change Eq. 5 to

2u
n

minimize =

T
i

Kui +

i=1

j j M 0

j=1

where M 0 is a constant. Since Eq. 6 is different only by a constant from Eq. 5, they have the same sensitivity and therefore
they will produce the same optimum solutions.
The strain energy and mass together will be minimized as much
as possible in Eq. 6, but if we do not need to reduce mass any
further after we meet a prescribed mass target M 0, we can change
Eq. 6 to

minimize =

i=1

1 T
u Kui + P
2 i

P = max 0,

iv i M 0

i=1

The function P satisfies the following three conditions: i


P is continuous, ii P 0 for all , and iii P = 0 if and
N
only if i=1
ivi M 0 0. With these properties of P, Eq. 7
can be changed to its mathematically equivalent form of a constrained optimization 8

2u
n

minimize =

T
i

Kui

i=1

subject to

v M
j j

j=1

It is important to note that Eq. 8 is the typical standard problem formulation for structural topology optimization with compliance minimization subject to a mass constraint.
From the equivalent form of topology optimization in Eq. 7,
sensitivity equations are derived into two different cases under the
assumption of linear elasticity and isotropy, as follows:
i

Case 1: Nj=1 j j M 0.
The sensitivity of the objective function with respect to
the density of the eth element is

=
e

2u
n

i=1

T K

ui + e

where K = Nj=1K j, and K j is the stiffness matrix of the jth


finite element see Ref. 22 for the details of this derivation. Assuming a constant density distribution within each
finite element in Eq. 9, K j can be written as
Journal of Biomechanical Engineering

10

where L j is the jth element stiffness matrix with unit density. The sensitivity then becomes

=
e

2
n

2
1

1 T
e u i K eu i

i=1

= ne

1
=
n i=1

= ne

Ua,e
+ e
e

1 T
u K eu i
2 i
e

Ua,e

e
n

2u
n

T K

i=1

+ e

Ua,e =

1
Ui,e
n i=1

ii Case 2: Nj=1 j j M 0.

=
e

u iTL eu i + e

i=1
n

1 T
u Kui + j j
2 i
j=1

i=1

K j = j L j

11

ui = ne

Ua,e
e

12

The bone-remodeling algorithm in Eq. 2 converges when the


right-hand side of the equation becomes zero, and the topology
optimization progress converges when the right-hand side of Eq.
11 or Eq. 12 becomes zero. When the mass constraint is not
satisfied Case 1, the form of the sensitivity equation the righthand side of Eq. 11 is exactly the same as that of the boneremodeling equation the right-hand side of Eq. 2. When the
mass constraint is satisfied Case 2, however, topology optimization becomes slightly different: The sensitivity equation, Eq. 12,
does not have a threshold value / n anymore, and the sensitivity
will be always negative as long as the mass constraint is met. It
means that there happens no bone osteoclast resorption regardless
of the magnitude of mechanical stimuli in a bone cell, whenever a
mass constraint is satisfied. From this comparison, we find that
topology optimization and the bone-remodeling algorithm will
have the same solutions when the mass constraint in topology
optimization is not met Case 1, but they will have slightly different solutions when the mass constraint is satisfied in topology
optimization Case 2.
By comparing Eq. 2 for bone remodeling and Eqs. 11 and
12 for topology optimization, we discover the mathematical relationships among B, n, , e, k, and :
B = ne

and

k=

13

Even though it is well known by experience that the SED-based


bone-remodeling algorithm and topology optimization produce
similar results, it has not been investigated how the constants in
the two approaches are mathematically related. This study indeed
for the first time revealed the relationships, as presented in Eq.
13. The multiplier B in the bone-remodeling algorithm is proportional to the number of load cases n, the exponent , and
the element volume e in topology optimization. We also find
that the reference SED k in the bone-remodeling algorithm is
proportional to the weighting factor for the mass constraint
but inversely proportional to the number of load cases n and the
exponent in topology optimization. With the relationship in
Eq. 13, the topology optimization and SED-based bone algorithm will theoretically produce the same solutions in Case 1 and
slightly different solutions in Case 2. Even though the global funcJANUARY 2009, Vol. 131 / 011012-3

Downloaded From: http://biomechanical.asmedigitalcollection.asme.org/ on 03/05/2015 Terms of Use: http://asme.org/terms

tional proposed by Harrigan and Hamilton 16 is mathematically


the same as the bone-remodeling equation, this study shows that
the typical topology optimization with a mass constraint and the
bone-remodeling equation are not always the same and hence will
not always produce the same solutions; this study examined the
analogy and differences between the two methods depending on
the mass condition.
Sensitivity filtering has been widely used to resolve numerical
instabilities in topology optimization. The original form, proposed
by Sigmund 21, is

g
=
k

1
N

14

j j

j=1
H
j

j=1

= r distk , j,
where g is a functional to be minimized, and H
j
min
j N distk , j rmin, k = 1 , . . . , N.
Incorporating the sensitivity filtering into Eq. 11, we have

=
e

j=1

jH
j
N

j=1

H
j

Ua,j
k =
j

G x,
N

Ua,j

j=1

15

We then realize that the topology optimization sensitivity equation with sensitivity filtering, Eq. 15, is very similar to the boneremodeling equation with a spatial influence function, Eq. 4, in
their form. The main difference is that Gix , in Eq. 15 is a
function of the density as well as the location.
Several researchers have modified the original sensitivity filtering equation, Eq. 14, to improve the performance: By using a
different type of weighting function, by moving a density term in
the denominator inside the summation 23, by dropping the density terms 24, and so on. In this paper, we used an exponential
= edistk,j/D instead of the linear weighting
weighting function H
j

function H j = rmin distk , j in order to avoid the difficulty of


choosing an equivalent filtering radius between different types of
weighting functions and dropped the density terms in Gix , in
Eq. 15 for equivalence.

Numerical Examples

4.1 Plate Model of Bone Tissue. For numerical comparison,


we first tested the plate model of bone tissue 5,6, which has been
frequently investigated in bone remodeling, as shown in Fig. 1.
The plate was meshed with 80 80 bilinear elements. The bone
material was assumed to be isotropic and linear elastic with
= 2 and c = 100 MPa g cm32 in Eq. 1. Poissons ratio was 0.3,
and the maximal density upper was 1.74 g cm3. For a sensor
influence parameter and a sensitivity filtering radius, D
= 0.025 mm was identically used in Eqs. 4 and 15. The initial
density distribution for bone-remodeling simulation and topology
optimization was = 0.8 g cm3 over the entire domain. We applied two load steps to simulate the transformation of morphology
under different loads. When equilibrium was reached in Load Step
1, an additional shear load was applied for Load Step 2. For terminating criterion, the total number of iterations in both load steps
was set 300.
For SED-based bone-remodeling, the constants in the boneremodeling equation, Eq. 2, were B = 1.0 g cm32
MPa time unit1 and k = 0.25 J g1 5,6. We used the forward
Euler integration scheme with a constant time step t = 0.0005 for
Eq. 4 to obtain the optimum results, as shown in Fig. 2a.
For topology optimization, we used the resultant mass of the
bone-remodeling simulation as the mass constraint M 0 in Eq. 8:
4.470 mg for Load Step 1 and 7.490 mg for Load Step 2. We

Fig. 1 Geometric and loading boundary condition for a plate


model of bone tissue: a Load step 1; b load step 2

performed a topology optimization with the aid of an optimality


criteria method 25. It is observed that the optimum results by

Fig. 2 Simulation result of structural configuration for a plate


model of bone tissue: a Strain energy density based bone
remodeling; b topology optimization

011012-4 / Vol. 131, JANUARY 2009

Downloaded From: http://biomechanical.asmedigitalcollection.asme.org/ on 03/05/2015 Terms of Use: http://asme.org/terms

Transactions of the ASME

Fig. 3 Geometric and loading boundary condition for a vertebral model

topology optimization Fig. 2b are structurally similar to the


results by the SED-based bone-remodeling algorithm Fig. 2a.
4.2 Vertebral Model. We used the vertebral model 26,27 as
the second numerical example, as shown in Fig. 3. The rectangular plate was meshed with 320 140 bilinear elements. The bone
material was assumed to be isotropic and linear elastic with
= 2 and c = 100 MPa g cm32 in Eq. 1. Poissons ratio was 0.3,
and the maximal density upper was 1.74 g cm3. For a sensor
influence parameter and a sensitivity filtering radius, D
= 0.025 mm was identically used in Eqs. 4 and 15. The initial
density distribution for bone-remodeling simulation and topology
optimization was = 0.8 g cm3 over the entire domain. For terminating criterion, the total number of iterations was set 400. The
physiological loading conditions 28 were considered using
bending moments and compressive load. In this paper, we used
three different load cases: right lateral bending with compressive
load, left lateral bending with compressive load, and pure compressive load. The moment for the left and right bending was
1.36 N m, this bending was implemented by applying a uniformly
distributed shear load on the top edge. In all cases, we applied a
uniformly distributed compressive load of 400 N on the top and
bottom edges 29.
The apparent strain energy density of the jth element was obtained by using weighted summation of the strain energy density
under each load case:
3

Ua,j =

cU
i

i,j

= c1U1,j + c2U2,j + c3U3,j

16

i=1

where subscript 1 denotes right bending, subscript 2 denotes left


bending, and subscript 3 denotes pure compression. The weighting factors used in this study were c1 = 0.1, c2 = 0.1, and c3 = 0.8.
For SED-based bone remodeling, the constants in Eq. 2 were
B = 1.0 g cm32 MPa time unit1, k = 0.20 J g1 26, and t
= 0.001 for the forward Euler integration scheme. Figure 4a
shows the optimum result.
With the given three load cases, the topology optimization
problem is stated as

c 2u
3

minimize =

T
i

Kui

i=1

subject to

v M
j j

17

j=1

The resultant mass of the bone-remodeling simulation was used


as the mass constraint for Eq. 17: M 0 = 2.6937 g. The same
Journal of Biomechanical Engineering

Fig. 4 Simulation result of structural configuration for a vertebral model: a Strain energy density based bone remodeling;
b topology optimization

weighting factors in Eq. 16 were used in Eq. 17. The result in


Fig. 4b is very similar to the structure obtained by the SEDbased bone-remodeling algorithm.

Discussion

We showed in Sec. 4 that the SED-based bone-remodeling algorithm and topology optimization produced very similar structures for two case studies. To quantitatively compare the progress
during both simulations, we examined the convergence histories
of mass and total strain energy during the simulations, as shown in
Figs. 5 and 6.
First, in terms of total strain energy, the final values in the 300th
iteration were very close, as shown in Figs. 5 and 6: The differences in the plate model of bone tissue were only 0.42% and
3.35% for Load Case 1 and Load Case 2, respectively, and 5.02%
in a vertebral problem.
Second, based on the experimental data from the literature, we
analyzed the mass convergence histories of the two approaches.
Xinghua et al. 30 examined the functional adaptation in long
bones of young and adult cavia porcellus to study the progress of
the bone-remodeling process quantitatively. They found that the
bone adaptation occurred mainly during the first 20 days: 65.30%
during the first 20 days, 26.30% during the second 20 days, and
8.40% during the third 20 days. Xin et al. 31 also studied the
adaptive bone remodeling in terms of the radius increase for adult
cavia porcellus: The radius increase in percentage was 75% during
the first month, 20% during the second month, and 5% during the
third month. In Figs. 5 and 6, we plotted the experimental data by
Xin et al. 31 for the comparison with numerical simulations in
this study. We find that in terms of the remodeling progress in the
time domain, the SED-based bone-remodeling algorithm produced results that are more consistent with the experimental results in the references; however, topology optimization showed
faster convergences.
The faster convergence of topology optimization in mass may
be explained by the priority of the optimization process: If the
initial design violates the mass constraint, satisfying the constraint
has priority over improving the objective function. Another aspect
JANUARY 2009, Vol. 131 / 011012-5

Downloaded From: http://biomechanical.asmedigitalcollection.asme.org/ on 03/05/2015 Terms of Use: http://asme.org/terms

Fig. 5 Convergence history during simulation of a plate model of bone tissue: a Mass change; b total strain energy
change

is that the SED-based bone-remodeling algorithm is formulated in


the time space, but topology optimization is directly derived in the
density space. That is, the SED-based bone-remodeling algorithm
gradually changes the bone density over the time domain; however, topology optimization undergoes a number of sudden big
changes in the density because the optimization algorithm directly
controls the density. It is observed that the topology optimization
is not proper to simulate the physiological remodeling process
obtaining intermediate states; instead, it produces the equivalent
final state more quickly and efficiently. In addition, if we use
topology optimization, we do not need to be concerned about
choosing a proper t, which plays a critical role in the Euler
integration scheme for stable progress in the bone-remodeling algorithm.
There are two main reasons that we obtained slightly different
results between topology optimization and SED-based boneremodeling equation. First, we have different sensitivities in Case
2 where the mass constraint is satisfied. In the plate problem in
Fig. 5a, it is observed that the mass constraint is satisfied with
some margin during the early iterations in Load Step 2 in the case
of topology optimization which is indicated by a few points with
densities smaller than 7.490 mg. This behavior is Case 2, which
is represented by Eq. 12, and during these iterations the topology
optimization progressed differently from the bone-remodeling

equation by Eq. 2. Second, the application of sensitivity values


can be different even in Case 1. In the case of topology optimization in this paper, we used the optimality criteria method for optimization. It adopts a heuristic update scheme based on theoretical sensitivity information by Eqs. 11 and 12. This update
scheme behaves slightly differently from the direct update algorithm in Eq. 2. In summary, the first factor is entirely based on
the difference in the mathematical formulation, and the second
factor is due to the choice of optimization or update method.
It should be addressed that the solutions by neither topology
optimization nor the bone-remodeling equation necessarily mean
the actual optimum design by the natureif we consider the
biological process as optimization. We considered only mechanical effects for the simulation, but there are other chemical and
biological phenomena that we did not consider in this study.
Moreover, even the mechanical requirements and properties in
bone are not fully understood yet. However, it is also interesting
to observe that the model with only mechanical parameters produce quite physiological results.
In this paper, we compared two different approaches for computational bone-remodeling simulationthe SED-based boneremodeling algorithm and structural topology optimizationfrom
the view point of mathematical formulations, supplementary numerical schemes, and behavior of the numerical solutions. From

011012-6 / Vol. 131, JANUARY 2009

Downloaded From: http://biomechanical.asmedigitalcollection.asme.org/ on 03/05/2015 Terms of Use: http://asme.org/terms

Transactions of the ASME

Fig. 6 Convergence history during simulation of a vertebral


problem: a Mass change; b total strain energy change

the comparison, we showed that the SED-based bone-remodeling


algorithm, which had been devised in biomechanics, can be mathematically matched with the structural topology optimization
problem. This study also revealed the mathematical relations
among the constants used in the SED-based bone-remodeling algorithm and topology optimization, and examined the analogy and
differences between the two methods depending on the mass condition; this relationship firmly shows that the bone indeed possesses the self-optimizing ability, as hypothesized in the field of
bone remodeling.

Acknowledgment
This work was supported by the Korea Research Foundation
Grant funded by the Korean Government MOEHRD, KRF2006-352-D00004.

References
1 Cowin, S. C., and Hegedus, D. H., 1976, Bone Remodeling I: Theory of
Adaptive Elasticity, J. Elast., 6, pp. 313326.
2 Fyhrie, D. P., and Carter, D. R., 1986, A Unifying Principle Relating Stress to

Journal of Biomechanical Engineering

Trabecular Bone Morphology, J. Orthop. Res., 4, pp. 304317.


3 Huiskes, R., Weinans, H., Grootenboer, H. J., Dalstra, M., Fudala, B., and
Sloof, T. J., 1987, Adaptive Bone Remodeling Theory Applied to Prosthetic
Design Analysis, J. Biomech., 20, pp. 11351150.
4 Beaupr, G. S., Orr, T. E., and Carter, D. R., 1990, An Approach for TimeDependent Bone Modeling and RemodelingTheoretical Development, J.
Orthop. Res., 8, pp. 651661.
5 Weinans, H., Huiskes, R., and Grootenboer, H. R., 1992, The Behavior of
Adaptive Bone-Remodeling Simulation Models, J. Biomech., 25, pp. 1425
1441.
6 Mullender, M. G., Huiskes, R., and Weinans, H., 1994, A Physiological Approach to the Simulation of Bone Remodeling as a Self-Organizational Control
Process, J. Biomech., 27, pp. 13891394.
7 Jacobs, C. R., Levenston, M. E., Beaupr, G. S., Simo, J. C., and Carter, D. R.,
1995, Numerical Instabilities in Bone Remodeling Simulations: The Advantages of a Node-Based Finite Element Approach, J. Biomech., 28, pp. 449
459.
8 Luenberger, D. G., 1989, Linear and nonlinear programming, 2nd ed.,
Addison-Wesley, Reading, MA.
9 Bendse, M. P., and Kikuchi, N., 1988, Generating Optimal Topologies in
Structural Design Using a Homogenization Method, Comput. Methods Appl.
Mech. Eng., 71, pp. 197224.
10 Bendse, M. P., 1989, Optimal Shape Design as a Material Distribution Problem, Struct. Optim., 1, pp. 193303.
11 Xie, Y. M., and Steven, G. P., 1993, A Simple Evolutionary Procedure for
Structural Optimization, Comput. Struct., 49, pp. 885896.
12 Hollister, S. J., Brennan, J. M., and Kikuchi, N., 1994, A Homogenization
Sampling Procedure for Calculating Trabecular Bone Effective Stiffness and
Tissur Level Stress, J. Biomech., 27, pp. 433444.
13 Fernandes, P. R., and Rodrigues, H. C., 1999, A Material Optimization Model
for Bone Remodeling Around Cementless Hip Stems, Proceedings of the 9th
European Conference on Computational Mechanics, Munich, Germany.
14 Bagge, M., 2000, A Model of Bone Adaptation as an Optimization Process,
J. Biomech., 33, pp. 13491357.
15 Huiskes, R., 2000, If Bone is the Answer, Then What is the Question? J.
Anat., 197, pp. 145156.
16 Harrigan, T. P., and Hamilton, J. J., 1994, Bone Remodeling and Structural
Optimization, J. Biomech., 27, pp. 323328.
17 Subbaraya, G., and Bartel, D. L., 2000, A Reconciliation of Local and Global
Models for Bone Remodeling Through Optimization Theory, ASME J. Biomech. Eng., 122, pp. 7276.
18 Jacobs, C. R., Levenston, M. E., Beaupr, G. S., and Simo, J. C., 1992, A
New Implementation of Finite Element-Based Remodeling, Proceedings of
the International Symposium on Computer Methods in Biomechanics & Biomedical Engineering, Swansea, pp. 57.
19 Diaz, A., and Sigmund, O., 1995, Checkerboard Patterns in Layout Optimization, Struct. Optim., 10, pp. 4045.
20 Matsui, K., and Terada, K., 2004, Continuous Approximation of Material
Distribution for Topology Optimization, Int. J. Numer. Methods Eng., 59, pp.
19251944.
21 Sigmund, O., 1994, Design of Material Structures Using Topology Optimization, Ph.D. thesis, Technical University of Denmark, Denmark.
22 Haug, E. J., Choi, K. K., and Komkov, V., 1986, Design Sensitivity Analysis of
Structural Systems, Academic, New York.
23 Borrvall, T., 2001, Topology Optimization of Elastic Continua Using Restriction, Arch. Comput. Methods Eng., 8, pp. 351385.
24 Sigmund, O., 2001, Design of Multiphysics Actuators Using Topology
OptimizationPart II: Two-Material Structures, Comput. Methods Appl.
Mech. Eng., 190, pp. 66056627.
25 Prager, W., 1968, Optimality Criteria in Structural Design, Proc. Natl. Acad.
Sci. U.S.A., 61, pp. 794796.
26 Xinghua, Z., He, G., Dong, Z., and Bingzhao, G., 2002, A Study of the Effect
of Non-Linearities in the Equation of Bone Remodeling, J. Biomech., 35, pp.
951960.
27 He, G., and Xinghua, Z., 2006, The Numerical Simulation of Osteophyte
Formation on the Edge of the Vertebral Body Using Quantitative Bone Remodeling Theory, Joint Bone Spine, 73, pp. 95101.
28 Wilke, H. J., Rohlmann, A., Neller, S., Schultheiss, M., Bergmann, G., Graichen, F., and Claes, L. E., 2001, Is it Possible to Simulate Physiologic Loading Conditions by Applying Pure Moments? A Comparison of In Vivo And In
Vitro Load Components in an Internal Fixator, Spine, 26, pp. 636642.
29 Tovar, A., Gano, S. E., Mason, J. J., and Renaud, J. E., 2005, Optimum
Design of an Interbody Implant for Lumbar Spine Fixation, Adv. Eng. Software, 36, pp. 634642.
30 Xinghua, Z., Xin, D., Yimin, Z., and Zhenyu, L., 1993, A Functional Adaptive Study in Long Bone, J. Biomed. Eng., 15, pp. 190192.
31 Xin, D., Xinghua, Z., and Xiuqin, Y., 1996, Adaptational Bone Remodeling
on the Radius After the Ulnar Osteotomy, J. Biomed. Eng., 15, pp. 190192.

JANUARY 2009, Vol. 131 / 011012-7

Downloaded From: http://biomechanical.asmedigitalcollection.asme.org/ on 03/05/2015 Terms of Use: http://asme.org/terms

Das könnte Ihnen auch gefallen