Journal Pre-proof A bone remodeling model governed by cellular micromechanics and physiologically based pharmacokinetics M.T. Bahia, M.B. Hecke, E.G.F. Mercuri, M.M. Pinheiro PII:
S1751-6161(19)31620-0
DOI:
https://doi.org/10.1016/j.jmbbm.2020.103657
Reference:
JMBBM 103657
To appear in:
Journal of the Mechanical Behavior of Biomedical Materials
Received Date: 23 October 2019 Revised Date:
11 January 2020
Accepted Date: 23 January 2020
Please cite this article as: Bahia, M.T., Hecke, M.B., Mercuri, E.G.F., Pinheiro, M.M., A bone remodeling model governed by cellular micromechanics and physiologically based pharmacokinetics, Journal of the Mechanical Behavior of Biomedical Materials (2020), doi: https://doi.org/10.1016/j.jmbbm.2020.103657. This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition of a cover page and metadata, and formatting for readability, but it is not yet the definitive version of record. This version will undergo additional copyediting, typesetting and review before it is published in its final form, but we are providing this version to give early visibility of the article. Please note that, during the production process, errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain. © 2020 Published by Elsevier Ltd.
A bone remodeling model governed by cellular micromechanics and physiologically based pharmacokinetics M. T. Bahiaa,b,∗, M. B. Heckeb , E. G. F. Mercurib,c , M. M. Pinheirod a
Departamento de Mecˆanica, Instituto Federal de Santa Catarina, Rua Pav˜ao, 1377, Costa e Silva, CEP 89220-618, Joinville, Brasil b Grupo de Bioengenharia, Programa de P´os-Graduac¸a˜ o em M´etodos Num´ericos em Engenharia, Universidade Federal do Paran´a, Curitiba, Brasil c Departamento de Engenharia Ambiental, Universidade Federal do Paran´a, Curitiba, Brasil d Departamento de Medicina, Universidade Federal de S˜ao Paulo, S˜ao Paulo, Brasil
Abstract This study describes a mathematical model for bone remodeling that integrates the bone cells activities with the pharmacological dynamics for bone-seeking agents. The evolution of bone cells population involves the osteoblast-osteoclast signaling mediated by biochemical factors and receives both mechanical stimulus evaluated at the microscale and pharmacological regulation. A physiologically based pharmacokinetic model (PBPK) for bone-seeking agents was developed to provide the drug concentration on bone sites and feed the remodeling algorithm. The drug effect on bone was reproduced coupling three different strategies: modification of the RANKL expression, increase the osteoclast apoptosis and change in the rate of differentiation of preosteoblasts. Computational simulations were performed in the PBPK model considering different dosing regimens. A 3D finite element model of a proximal femur was generated and the simulation of the ∗
Corresponding author Email address:
[email protected] (M. T. Bahia)
Preprint submitted to Journal of the Mechanical Behavior of Biomedical MaterialsJanuary 30, 2020
bone remodeling algorithm were implemented in Matlab. The results indicate that the proposed integrated model is able to capture adequately the expected adaptive behavior of bone subjected to mechanical and pharmacological stimulus. The model demonstrated to have potential for use as a platform to investigate therapies and may help in the study of new drugs for bone diseases. Keywords: bone cell dynamics, physiologically based pharmacokinetics (PBPK), continuum michomechanics, osteoporosis, mathematical modeling
1
1. Introduction
2
Bones are a living tissue capable of adapting mass and microstructure in re-
3
sponse to mechanical and biochemical stimuli in a physiological process called
4
bone remodeling. At the cellular level, remodeling involves primarily osteoblast-
5
mediated bone formation and osteoclast-mediated bone resorption. Osteocytes
6
acts as mechanotransducers receiving mechanical stimulus and producing bio-
7
chemical signaling agents (Fig. 1), such as nitric oxide (ON) and prostaglandin
8
E2 (PGE2 ), that will drive osteoblasts and osteoclast response (Rieger, 2011; Ross
9
et al., 2017; Qin and Wang, 2012). Bone diseases such as osteoporosis, osteope-
10
nia and Paget’s disease (Barkaoui, 2012) cause imbalances in bone remodeling.
11
In the case of osteoporosis, there is a degradation of the bone architecture and
12
structural properties that results in decreased bone quality and greater risk of
13
fractures (NIH, 2001).
14
The mortality rate among patients one year after osteoporotic fractures is
15
15 to 30 % due to complications such as infections, venous thrombosis, pressure
16
ulcers, or comorbid conditions, particularly cardiovascular diseases (Pinheiro,
17
2008; Pinheiro and Eis, 2010).According to Fortes et al. (2008) only 30 % of these 2
18
patients return to their daily activities and 11.6 % become totally dependent. In
19
Brazil, it is estimated that the cost of hospitalization due to osteoporotic fractures
20
in a private hospital is approximately US$ 12, 000 per patient (over 50 years of
21
age) (Ara´ujo et al., 2005). Osteoporosis represents a public health concern given
22
its high economic and social impact. With the increase in life expectancy, its
23
burden tends to increase.
Figure 1: Cartoon representing the bone remodeling process with the osteoblast-mediated bone formation and the osteoclast-mediated bone resorption driven by osteocytes. Osteocytes act in remodeling similarly to a master builder or a regent in an orchestra, signaling for osteoblasts and osteoclasts according to the mechanical/biochemical stimuli received. The cartoon shows a sleeping osteocyte, which is intended as a humorous content, as it is highly unlikely that an osteocyte would be slacking off amidst all the multiple signals it receives, be it direct strain, hydrostatic pressure, or interstitial fluid flow (Pastrama et al., 2018; Scheiner et al., 2016).Art: Miguel T. Bahia.
24
Given its financial and social relevance, the understanding of the processes
25
that lead to bone diseases and possibly to the development of new efficient ther3
26
apies are of paramount importance. Mathematical models of bone remodeling
27
have the ability to create virtual scenarios that allow the study of the dynamics
28
of bone cells populations, drug effects and different therapeutic strategies.
29
The approaches adopted to simulate bone remodeling numerically can be
30
quite different. Among them are mechanistics, phenomenologicals, multiscale,
31
and optimization methods, and hybrid techniques (Rieger, 2011; Doblar´e and
32
Garcıa, 2002; Garc´ıa-Aznar et al., 2005; Ganghoffer, 2012; Goda et al., 2016; Louna
33
et al., 2019). Mathematical models that consider the evolution of bone cells pop-
34
ulations or cell activities, include the pioneering works of Kroll (2000), which
35
analyzes the dual action of the parathyroid hormone (PTH) on bone formation
36
and resorption; Komarova et al. (2003), which models the autocrine and paracrine
37
interactions of osteoblasts and osteoclasts; and the model of Lemaire et al. (2004)
38
that explicitly expresses the RANK-RANKL-OPG pathway. Based on Lemaire
39
et al. (2004) model, the work of Pivonka et al. (2008) rewrites the signaling ex-
40
pressions with activation and repression functions from chemical kinetics (Hill
41
functions) and incorporates other signaling into the model. Another important
42
work is Scheiner et al. (2013) that combines the Pivonka et al. (2008) model with
43
the continuum micromechanics and incorporates the mechanical stimulus evalu-
44
ated in the microscale and use the Mori-Tanaka homogenization technique (Mori
45
and Tanaka, 1973).
46
Numerical simulations with pharmacological models represent a whole new
47
avenue of possibilities for bone remodeling studies. Parameters such as dose, fre-
48
quency, route of administration and effectiveness of a medicine can be included
49
in the model, making the simulation quite relevant in terms of treatment, preven-
50
tion and cure of bone diseases. The application of the pharmacological modeling 4
51
involves (i) the development of a pharmacokinetic model that simulates the tem-
52
poral evolution of the drug concentration in the body or a specific organ, (ii) the
53
development of a pharmacodynamic model that simulates the drug’s action in
54
the body or organ. In other words, how the drugs interacts with the body, and
55
how the body reacts to the drug. In both cases, the models should be able to
56
reasonably simplify a complex physiological system into a tractable form.
57
Compartment models are the most adopted approach for pharmacokinetics
58
modeling and allow the combination of dose and individual processes of absorp-
59
tion, distribution, metabolism and excretion (ADME) in a logical and straightfor-
60
ward manner (Rosenbaum, 2016). In this sense, the body is viewed as a system of
61
imaginary and homogeneous compartments, with no anatomical or physiologi-
62
cal representation and where drug concentration is uniform. Each compartment
63
correspond to groups of tissues with similar drug distribution rates, described by
64
mass balance equations.
65
In general, constant kinetic rates and other parameters are estimated using
66
curve fitting techniques from concentration-time profiles of drugs in the blood,
67
tissue, or other easily accessible body fluids such as urine, faeces or breast milk
68
(McMullin, 2005). The adjustment parameters obtained are mathematically rele-
69
vant, but they lack a physical meaning, which can limit the model as a predictive
70
tool, particularly for interspecies extrapolation. Another drawback of the model
71
is the difficulty, or even impossibility, to establish a straightforward relation-
72
ship between blood concentration-time profile and the intended site of action
73
concentration-time profile (Nestorov, 2003).
74
An alternative to the classical compartment models are the physiologically
75
based pharmacokinetics (PBPK) models. They also represent tissue or organs as 5
76
compartments, but describe the interconnected body compartments with perti-
77
nent physiological, biochemical and physicochemical information. Furthermore,
78
the route of drug administration respect their proper relationship with the over-
79
all physiology. Such features makes PBPK models a suitable tool for all types
80
of pharmacokinetic extrapolations, allowing the prediction of events in humans
81
from extrapolations of pre-clinical in vitro data or through intra-species extrapo-
82
lations from clinical data (Nestorov et al., 1998). A review of the PBPK modeling
83
approaches for drug research and analysis of advantages and limitations can be
84
found in Nestorov (2003).
85
The combination of dynamical modeling of cellular populations with pharma-
86
cological modeling has been used in a few studies. Among those are the works of
87
Scheiner et al. (2014) and Hambli et al. (2016) that incorporate a pharmacokinetic
88
(PK) model of the drug Denosumab, used in osteoporosis treatment, coupled to
89
the models of Pivonka et al. (2008) and Komarova et al. (2003), respectively, as a
90
pharmacodynamics model for quantifying the effect of denosumab on bone re-
91
modeling. Both works used a classical two compartment model, one for the sub-
92
cutaneous tissue and another for the blood serum. The results of Hambli et al.
93
(2016) showed an increase of BMD in the femoral neck, trochanter and total hip
94
compared with placebo in conformity with the results reported in the literature.
95
The model of Scheiner et al. (2014) allows reasonable simulations of osteoporotic
96
scenarios and pharmacological intervention through the administration of the
97
drug denosumab.
98
This research aims to integrate a physiologically-based pharmacokinetic model
99
(PBPK) with a pharmacodynamic model for bone remodeling based on the dy-
100
namics of bone cells populations. The modeling of cells populations describes 6
101
the signaling between osteoblasts and osteoclasts based on the model of Lemaire
102
et al. (2004), with the addition of the mechanical stimulus proposed by Scheiner
103
et al. (2013) and a pharmacological stimulus that will drive the bone remodeling
104
process. The PBPK model provides the drug concentration in the bone, which
105
allows the computation of a regulatory function that influences the dynamics
106
of bone cells population. The drug’s antiresorptive action is represented here
107
by a loss of efficacy of the active osteoclasts. We simulate this effect by adding
108
a regulatory function in the RANKL expression and increasing the osteoclasts
109
apoptosis rate. Therefore, the proposed model combines the advantages of phys-
110
iologically based pharmacokinetic models, whose compartments correspond to
111
exact anatomical or physiological entities, with a cellular population model that
112
better captures the complex osteoblast-osteoclast interactions in the microme-
113
chanical environment.
114
2. Materials and methods
115
This work proposes an unified model for bone remodeling, coupling the dy-
116
namics of bone cell populations and a pharmacokinetic model. In summary, the
117
procedure involves a finite element analysis performed in a three-dimensional
118
proximal femur. Geometry and material properties were modeled based on CT
119
images from previous work (Bahia et al., 2019). Once the structural analysis is
120
available, two stimuli to the bone cells are inputed to the remodeling process -
121
(i) a mechanical stimulus evaluated at the microscale, i.e. the microstrain energy
122
density; and (ii) a pharmacological stimulus, function of the drug concentration
123
on the bone. The strain energy density evaluated at the microscale is obtained
124
by a localization procedure (Mori-Tanaka scheme) performed in the nonhomo7
125
geneous bone medium. The concentration of the drug in the bone is provided by
126
a pharmacological model based on physiology, as can be seen in Fig. 2.
Figure 2: Overview of the bone remodeling procedure integrating the dynamics of bone cells populations and a physiologically based pharmacokinetics model. Mechanical loadings and local drug concentration are the stimuli in the bone remodeling process model. A micromechanical representation is used for the homogenization and scaling down of the strain energy densities for an extravascular bone matrix level. Art: Prepared by the author.
127
2.1. Physiologically based pharmacokinetic model
128
A physiologically based pharmacokinetic model describes the drug concen-
129
trations in the blood and other organs using physiological parameters such as
130
blood flow and organ volume as well as compartments representing different tis-
131
sues and organs. Here the PBPK model will provide the drug concentration at
132
every bone site, and is converted into a pharmacological stimuli in the dynamics 8
133
of bone cells populations model, represented by a system of ordinary differential
134
equations.
135
A schematic of the PBPK model structure is given in Fig. 3. As can be seen
136
in the model, the body is divided into a number of tissue compartments, each
137
one characterized by appropriate volume and flow rates. For simplicity, here
138
we adopt the following compartments: blood, gut, liver, kidney, rapid perfused
139
tissues, slowly perfused tissues and bone.
140
The temporal evolution of the drug concentration in each tissue/organ is de-
141
scribed with a mass balance equation that accounts for tissue blood flow rate
142
(Qi ) and the concentration difference between the arterial blood (cA ) and the ve-
143
nous blood (cVi ). Each tissue bone flow rate is assumed to be proportional to the
144
cardiac output (Q) and based on reference values for a 70 kg man (ICRP, 1975;
145
Yokley et al., 2006). Mean cardiac output is assumed to be 276 liters per hour.
146
Physiological values are adjusted to satisfy the requirement Q = ΣQi , where
147
i correspond to the tissues or organs: gut, kidney, liver, rapid perfused tissues,
148
slowly perfused tissues and bone. The volumes of each organ are considered
149
fractions of the body weight (BW) in [kg] as seen in Tab. 1.
150
The route of administration for a drug may be intravenous, or oral (gastroin-
151
testinal tract). In addition to these routes of administration, there are also pul-
152
monary (inhalation), subcutaneous (adhesives), muscular (injection) and oph-
153
talmic (eye drops). Drugs administered orally may be absorbed in different parts
154
of the gastrointestinal tract. Drugs absorbed through the oral cavity or rectum
155
bypass the liver. All other parts of the alimentary canal (especially the stom-
156
ach, duodenum, small and large intestines) carry blood (and drugs) to the liver
157
through the portal vein. There are also transporters in the mucosal cells of the gut 9
Figure 3: Physiologically based pharmacokinetic (PBPK) model. Qi terms are blood flow rates, c terms are concentrations. Arterial input is on the right and venous effluent in on the left. Source: Prepared by the author.
158
wall that pick up the drug from the gut and may excrete the drug back into the
159
intestine. Relatively high drug concentrations are processed by the liver during
160
the first passage of the drug. The liver can excrete the drug in the bile and then
161
into the duodenum. Therefore, the drug can be reabsorbed in an enterohepatic
162
cycle (Gieschke and Serafin, 2013).
163
The main organs for drug elimination are the liver and the kidneys. The liver
164
with its enzymatic equipment is the main metabolizing organ in the body. The
165
kidneys filter out wastes (including small drugs) from the blood and excrete them 10
166
with urine. Drug absorption by tissues depend on blood perfusion and tissue
167
permeability.
168
In the case of drugs used for treatment and prevention of osteoporosis, focus
169
of this investigation, they can be divided basically into two categories : anti-
170
resorptive agents or anabolic agents. Anti-resorptive agents reduce bone resorp-
171
tion, leading to increased BMD by several degrees. Estrogen, raloxifene, bispho-
172
sphonates and the human monoclonal antibody to RANKL are included in this
173
category. On the other hand, anabolic agents such as teriparatide (PTH1-34) and
174
total parathyroid hormone (PTH1-84), stimulate bone formation, thus increasing
175
BMD (Szulc and Bouxsein, 2011).
176
Most of the pharmacokinetic characteristics of bisphosphonates have been
177
derived from urinary excretion data since the plasma concentration following
178
therapeutic doses generally falls below the sensitivity limits of the quantification
179
assays (Porras et al., 1999). The bisphosphonate is eliminated from the plasma
180
by bone deposition and urinary excretion. For illustrative purposes, in rats it has
181
been found that 30 to 40% of 1mg/kg alendronate is eliminated in urine within 24
182
hours after dosing (Lin et al., 1992). About 60 to 70% of the dose is sequestrated to
183
the bone with short-term effect. In rats, the rate of renal excretion of alendronate
184
reached a maximum of about 25 mg/min/kg (Lin et al., 1991).
185
The mass balance equations for the PBPK model illustrated in Fig. (3) are
186
listed in Eqs. (1). The variables Vi , ci , Qi , t are volume, concentration and blood
187
flow in the tissue/organ i, and time, respectively. The indices i refers to blood, gut,
188
liver, kidney, rapidly perfused tissues (rp), slowly perfused tissues (sp) and bone,
189
respectively. Note that the variables Q, cV and cA represent the cardiac output
190
(blood flow), concentration in venous flow and concentration in arterial flow, 11
191
respectively. Indices A and V in each concentration expression refers to arterial
192
affluent and venous effluent, respectively, of an especific organ. For instance,
193
cA is the affluent concentration entering in each organ and cVbone is the effluent
194
concentration leaving the bone. Tab. (1) has the description and values of all the
195
symbols used in Eqs. 1a-1g.
dcblood = Q (cV − cA ) dt dcgut Vgut = Qgut cA − cVgut dt dcliver Vliver = Qliver cA + Qgut cVgut − (Qliver + Qgut ) cVliver dt
Vblood
− CLhepato−bilis cVliver Fu = Qkidney cA − cVkidney − CLrenal cVkidney Fu
dckidney dt dcrp Vrp = Qrp cA − cVrp dt dcsp = Qsp cA − cVsp Vsp dt dcbone = Qbone (cA − cVbone ) Vbone dt
Vkidney
(1a) (1b)
(1c) (1d) (1e) (1f) (1g)
196
The concentrations at t = 0 are cblood = cgut = cliver = ckidney = crp =
197
csp = cbone = 0. Furthermore, cA = cb , cVgut = cgut /Rgut , cVliver = cliver /Rliver ,
198
cVkidney = ckidney /Rkidney , cVrp = crp /Rrp , cVsp = csp /Rsp , e cVbone = cbone /Rbone .
199
Indices A and V in each concentration expressions refers to arterial and venous
200
effluent, respectively. Thus, the total venous concentrations is given by:
cV =
(Qliver + Qgut ) cVliver + Qkidney cVkidney + Qrp cVrp + Qsp cVsp + Qbone cVbone Q (2) 12
201
Finally the total blood flow is expressed by: Q = Qliver + Qgut + Qkidney + Qrp + Qsp + Qbone
(3)
202
The proposed PBPK model was fully written in MatLab 7.12 (R2011a) software
203
(MathWorks, Natick, Massachusetts, US) and run in a standard personal com-
204
puter (Intel Core i7 5500U CPU 2.4 GHz). The physiological parameters adopted
205
here combine data or are derived from different sources, such as O’Flaherty
206
(1991b); Pertinez et al. (2013) and are listed in Tab. (1).
207
2.1.1. Bone tissue modelling
208
The mineralized components of bone tissue are described by an adaptation
209
of the permeability rate limited tissue compartment model (Nestorov, 2003; Per-
210
tinez et al., 2013). According to this model, the bone is composed of a vascular
211
compartment, defined as the bone surface, and an extravascular compartment,
212
defined as the bone matrix. The bone remodelling process will govern the ex-
213
change between these two compartments. In fact, whereas bone is woven into
214
two main patterns (cortical and trabecular), each with different appearance and
215
characteristics, the bone tissue model presents a total of two compartments both
216
with two sub-compartments: cortical bone matrix, cortical bone surface, trabec-
217
ular bone matrix and trabecular bone surface. Expressing in mathematical terms,
218
the temporal evolution of the drug concentration in the vascularized bone sur-
219
face cvasi and in the extravascular bone matrix cbmi with index i assuming cortical
220
or trabecular types, can be modeled as: dcvasi Qbone cA Qbone cvasi cvasi cbmi = − Rbone − kform + kres dt Vvasi Vvasi Vvasi Vbmi dcbmi cvasi cbmi = kform − kres dt Vvasi Vbmi 13
(4) (5)
Table 1: Physiological parameters used in the PBPK model.
Simbol
Units
Value
Description
Q
l/h
276
Cardiac output
Qgut
l/h
0.19Q
Blood flow to guta
Qkidney
l/h
0.19Q
Blood flow to kidneya
Qliver
l/h
0.06Q
Blood flow to livera
Qbone
l/h
0.19Q
Blood flow to bonea
Qrp
l/h
0.36Q
Blood flow to well perfused tissuesa
Qsp
l/h
0.125Q
Blood flow to slowly/poorly perfused tissues
BW
kg
70
Body weighte
Vgut
l
0.2895BW
Gut volumeb
Vkidney
l
0.004BW
Liver volumec
Vliver
l
0.025BW
Kidney volumec
Vrp
l
0.04BW
Well perfused tissues volumec
Vbone
l
0.026BW
Bone volumed
Vsp
l
4
Poorly perfused tissues volume
Rgut
n/a
0.9
Partition coefficient gut/plasmab
Rliver
n/a
0.47
Partition coefficient liver/plasmab
Rkidney
n/a
3
Partition coefficient kidney/plasmab
Rrp
n/a
0.9
Partition coefficient well perf. tissues/plasmab
Rsp
n/a
0.6
Partition coefficient poorly perf. tissues/plasmab
Rbone
n/a
2
Partition coefficient bone/plasma
CLrenal
n/a
0.03
Renal clearanceb
CLhepato bilis
n/a
0.03
Hepato-biliary clearanceb
Fu
n/a
0.73
Fraction unbound in blood b
a
Brown et al. (1997).
b
Pertinez et al. (2013).
c
Yokley et al. (2006).
d
O’Flaherty (1991b).
e
ICRP (1975).
14
221
where the indexes vas and bm refer to vascularized bone surface and extravas-
222
cular bone matrix, respectively. The terms kform and kres describe the bone for-
223
mation and bone resorption rates, respectively, and are both defined in the bone
224
cells populations model. Fig. 4 illustrates the bone tissue compartment model.
225
The steady-state values of total vascularized bone surface cvas and the total
226
extravascular bone matrix cbm will drive the dynamics of bone cells populations.
227
These total values are obtained by adding the cortical and trabecular contribu-
228
tions as follows:
229
cvas = cvascort + cvastrab
(6)
cbm = cbmcort + cbmtrab
(7)
where the indexes cort and trab refer to cortical and trabecular types, respectively.
Figure 4: Bone tissue compartment model. Source: Prepared by the author. 230 231
Bone physiological parameters such as the total bone volume, and the specific
232
surface and matrix volume fractions for the cortical and trabecular compartments 15
233
were extracted or derived from O’Flaherty (1991a) following the procedure of
234
Pertinez et al. (2013). The bone weight was calculated as 0.564 times total marrow
235
free dry bone weight considering a density of 1.92 g/ml for the marrow free dry
236
bone.
237
2.2. Cells populations model
238
Bone remodeling is a biological process of bone adaptation that involves se-
239
quential osteoblast-mediated bone formation and osteoclast-mediated bone re-
240
sorption at the same location, acting to maintain bone homeostasis, repairing
241
small defects and replacing bone matrix (Burr and Allen, 2013). A complex net-
242
work regulates the interactions between bone cells. The signaling pathway RANK-
243
RANKL-OPG has been shown to be a major regulator of bone remodeling act-
244
ing on osteoclastogenesis (Hofbauer et al., 2004; Martin, 2004). RANKL is a cy-
245
tokine (membrane bound factor) produced by osteoblasts and stromal cells that
246
binds to its receptor RANK expressed by osteoclasts, and induces both differ-
247
entiation and activation of osteoclasts. Osteoprotegerin (OPG), also produced
248
by stromal/osteoblastic cells, binds to RANKL to prevent RANKL stimulation by
249
RANK, inhibiting bone resorption (Pivonka et al., 2008).
250
Imbalances in the RANK-RANKL-OPG pathway are closely associated with
251
diseases such as osteoporosis, osteopenia, and tumors, among others. It is spec-
252
ulated that an overproduction of RANKL is associated with the development
253
of postmenopausal osteoporosis (Miller, 2009). Furthermore, the transforming
254
growth factor (TGF-β) stimulates osteoblastic differentiation and promotes bone
255
formation and hormones like the parathyroid hormone (PTH) acts by inhibiting
256
OPG and enhancing RANKL expression.
257
In this context, mathematical modeling provides a powerful tool to study the 16
258
interactions between bone cells, their biochemical regulators, allowing the un-
259
derstanding of the bone remodeling process in distinct scenarios. Lemaire et al.
260
(2004) proposes a mathematical model for the temporal evolution of the bone
261
cell populations through a system of differential equations. The model includes
262
interesting features like the osteoblast-osteoclast coupling, the RANK-RANKL-
263
OPG pathway, and the antagonistic effects of PTH administration on the control
264
of OPG and RANKL expression by active osteoblasts. These features enabled
265
the assessment of certain bone metabolic dysfunctions. In addition, the model
266
includes the effect of growth factor TGF-β on the differentiation of precursor
267
osteoblasts.
268
Here, some modifications were applied to the original model by Lemaire et al.
269
(2004) to incorporate a mechanical stimulus based on strain energy density eval-
270
uated at the microscale (Scheiner et al., 2013) and the pharmacological regulation
271
of an anti-resorptive drug on the bone matrix (Ross et al., 2017). The dynamics
272
of the proposed bone cells populations obtained, built upon the work of Lemaire
273
et al. (2004) are defined in Eqs. 8-18. The variables OBp , OBa and OCa refer to
274
pre-osteoblasts, active osteoblasts and active osteoclasts, respectively. BA and
275
BI are the antiresorptive drug concentrations in the bone matrix and plasma, re-
276
spectively. The behavior of bone cells is regulated by the binding of cytokines to
277
the receptors on the cells surface. The chemical kinetics of reactions can be rep-
278
resented by the law of mass action. A graphical scheme of the model is presented
279
in Fig. 5.
17
Figure 5: Sketch of the bone cells populations model. Microscopic strain energy density (microSED) and drug concentration plays a regulatory role in the dynamics of cells populations. These stimuli are added to the previous work of Lemaire et al. (2004). Artwork: Miguel T. Bahia.
dOBp dt dOBa dt dOCa dt dBA dt
= DR πC + POBp OBp Πε − =
DB OBp πC
DB OBp − kB OBa πC
(8) (9)
= DC πL − (1 + k7 BA)DA πC OCa
(10)
= k8 cdrug (1 − BA) − k9 BA
(11)
280
where cdrug is the local drug concentration that reaches the bone obtained from
281
the PBPK model. For simplicity, we adopted the steady-state value of the total
282
extravascular bone matrix, i.e., cdrug = cbm . BA is the drug amount that binds
283
to the bone matrix. DB is a factor of proportionality and πC is the influence of
284
TGF-β, expressed by: DB = f0 dB πC = 18
C + f0 C S C + CS
(12)
285
RANKL induces the differentiation of active osteoclasts which release TGF-β
286
causing osteoclastic apoptosis and controlling osteoblast differentiation.
287
The drug concentration in the bone obtained from the PBPK model is used
288
here to trigger an antiresorptive effect through a pharmacoregulation function
289
ΠBA .
290
291
The RANKL influence, πL is obtained by: IL + PRLεbm k3 KLP .πP .OBa P 1 + ΠBA + πL = KO k4 1 + k3 + k1 rL k4 k2 k0 πP +I0 where πP is the fraction of occupied PTH receptors which is expressed by: πP =
292
293
(13)
IP kP IP kP
+ +
SP kP k6 k5
(14)
and the term PRLεbm corresponds to the variation of RANKL related to changes in strain energy. It is activated only when in disuse and is expressed by: ! wεbm PRLεbm = κ 1 − wεbmst
(15)
294
where the adjustment constant κ is null for wεbm ≥ wεbmst . and positive other-
295
wise. The incorporation of this non-null term mathematically expresses that any
296
deformation would limit the external production of RANKL by osteocytes and
297
cap marrow stromal cells, and would be totally stopped if the critical value of
298
SED is reached. Physiologically, the mechanotransductive activity of osteocytes
299
located in lacunar pores of the extravascular bone matrix would be signaled in at
300
least three ways: direct deformation of osteocytes, hydrostatic pressure or even
301
the presence of fluid flow. As Scheiner et al. (2013) points out, all of these sit-
302
uations are associated with a decrease in the RANKL/ OPG rate expressed by
303
osteocytes. 19
304
The term ΠBA is a drug-regulatory function that simulates the effect of the
305
external administration of an antiresorptive drug on the population of cells. The
306
term was added to the original expression proposed by Lemaire et al. (2004) for
307
the RANKL concentration equation, being expressed by: ΠBA = kBA BA
308
(16)
where BA is given by Eq. 11 and kBF is an adjustment constant.
309
The population of preosteoblasts OBp increases with the differentiation of
310
uncommitted progenitors osteoblast cells with a maximal differentiation rate DR ,
311
promoted by the growth factor TGF-β.
312
Uncommitted progenitor osteoblasts represent the initial stage of maturation
313
of the osteoblast lineage, represented by a constant pool of mesenchymal stem
314
cells located in the bone marrow, spleen and liver (Gowder, 2012). As pointed
315
out by Lemaire et al. (2004), this category corresponds to a composite of several
316
phenotypes that share similar characteristics.
317
On the other hand, the population of precursor osteoblasts decreases due
318
to the differentiation of the precursor osteoblasts into active osteoblasts which
319
occurs at a maximum rate.
320
Following the proposal of Scheiner et al. (2013), the mechanical stimulus is
321
modeled adding a term in Eq. 8 that regulates the proliferation of osteoblasts
322
as a function of the strain energy density evaluated locally. The function Πε
323
is the mechanoregulation and PR is the constant indicating the proliferation of
324
osteoblasts according to the strain energy density:
Πε = Πεst
1+λ 20
!! wεbm −1 wεbmst
(17)
325
where Πεst is the equilibrium value of Πε , wεbm is the value of local strain en-
326
ergy density and wεbmst is the equilibrium value. Parameter λ is an adjustment
327
constant that is null to wεbm ≤ wεbmst and positive otherwise. In the presence of
328
bone diseases such as osteoporosis or even during a treatment with antiresorp-
329
tive agents this parameter should be recalibrated since it significantly affects the
330
bone response to mechanical stimuli. Current studies suggest that precursor os-
331
teoblasts may respond to mechanical stimuli, either matrix deformations and/or
332
fluid flow, and osteocyte-derived biochemical signaling (i.e. sclerostin). In the
333
present work, a phenomenological activating function was adopted, following
334
the proposal of Scheiner et al. (2013), as seen in Eq. 17. As shown in Eqs. ? and
335
?, the maximum proliferation rate of precursor osteoblasts is reached when the
336
mechanoregulator term equals to 1, which would provide sufficient mechanical
337
activation of osteoblasts. According to the studies of Jones et al. (1991) and Kas-
338
par et al. (2002), low straining levels would reduce the proliferation rate by 25 and
339
50%, which is simulated by the adoption of a minimum value of the mechanoreg-
340
ulatory function Πεst related to the limit value wεbm . Above this value the in-
341
creased proliferation of precursor osteoblasts is activated. Thus, the function
342
provides intermediate values between the maximum and minimum proliferation
343
rates.
344
The model’s dead zone is defined in the mechanical and pharmacological
345
stimuli expressions. In the case of mechanical stimulation, Eq. 17 provides null
346
values when the strain energy density is within the corresponding range. When
347
the drug concentration is zero, the expression of the pharmacological stimulus
348
produces null values.
349
The active osteoblast population OBa increases according to the pre-osteoblast 21
Table 2: Parameters used in the simulation of the bone remodeling model. Values are extracted from Lemaire et al. (2004), except for ki (i = 7, ..., 12) and kBA , that were adjusted by the authors Symbol
Units
Value
Description
5 × 10−3
Number of OC to get half differentiation flux
DA
day
−1
0.7
Rate of osteoclast apoptosis caused by TGF−β
dB
day−1
0.7
Cs
pM
Differentiation rate of responding osteoblasts
DC
pM.day
−1
2.1 × 10
Differentiation rate of osteoclast precursors
DR
pM.day−1
7 × 10−4
Differentiation rate of osteoblast progenitors
−3
f0
n/a
0.05
Fixed proportion
IL
pM.day−1
0 − 106
Rate of administration of RANKL
IO
pM.day−1
0 − 106
Rate of administration of OPG
IP
pM.day
6
0 − 10
Rate of administration of PTH
10
Fixed concentration of RANK
10−2
Rate of OPG-RANKL binding
−1
K
pM
k1
pM−1 .day−1
10
Rate of OPG-RANKL unbinding
k3
pM−1 .day−1
5.8 × 10−4
Rate of RANK-RANKL binding
k4
pM.day−1
1.7 × 10−2
Rate of RANK-RANKL unbinding
k2
day
−1
0.02
Rate of PTH binding with its receptor
k6
day−1
3
Rate of PTH unbinding
kB
day−1
0.189
KLP
pM/pM
kO
day−1
KOP
day
2 × 10
Minimal rate of production of OPG per cell
kP
day−1
86
Rate of elimination of PTH
rL
pM.day−1
103
Rate of RANKL production and elimination
SP
pM.day
k5
pM .day −1
−1
−1
Rate of elimination of active osteoblasts
3 × 10
Maximum RANKL attached on each cell surface
0.35
Rate of elimination of OPG
6
5
250
Rate of synthesis of systemic PTH
λ
n/a
1.2
Parameter of anabolic adjustment
κ
n/a
5 × 102
Parameter of RANKL inibition
Πεst
n/a
0.5
Equilibrium of mechanoregulation function
wεbmst
Pa
0.78
SED inferior limit
k7
n/a
0.3
−1
Hill function parameter for apoptosis of OCa
6 × 10
Bisphosphonate transfer from plasma to bone
k9
1/h
5 × 10−4
Bisphosphonate on bone clearance rate
k10
1/h
25 × 10−5
First order BMD rate OCa /OBa
k11
1/h
1.5 × 10
BMD balance
k12
n/a
0, 7979
Baseline value of the OCa /OBa
kBA
n/a
1
Parameter of the Hill function for drug regulation
k8
pM.day
−1
−3
−4
22
350
differentiation, which occurs with the maximum differentiation rate DB and is
351
inhibited by the TGF-β. The population is reduced by the apoptosis of active
352
osteoblasts, which follows the apoptosis rate kB .
353
Antiresorptive drugs such as bisphosphonates bind to the hydroxyapatite in
354
the bone and inhibit resorption by osteoclasts. In the model, the loss of func-
355
tion of osteoclasts is simulated by a change in RANKL concentration in Eq. 13
356
and by an increase in the rate of apoptosis of active osteoclasts in Eq. 10. The
357
bisphosphonate attaches to the bone in a second-order forward, first-order re-
358
verse reaction with hydroxyapatite attachment sites (Ross et al., 2017), resulting
359
in values in the range [0, 1]. In Eq. 11, the drug concentration in the bone ma-
360
trix BA follows Ross’s proposal. The plasmatic anti-resorptive agent BI attaches
361
to hydroxyapatite cristals in bone with a plasma-to-bone transfer rate k8 and is
362
eliminated at a rate k9 .
363
The model proposed by Ross et al. (2017) incorporates a reference biomarker
364
for the bone mineral density (BMD) into the system of differential equations (Eqs.
365
8-11), here denoted as the dimensionless variable BM in the Eq. 18. The evolution
366
of BMD is governed by the difference between the osteoblast/osteoclast ratio and
367
its reference value, and by the difference between the biomarker and a baseline
368
value: dBM = k10 dt
OBa − k12 − k11 (BM − 1) OCa
(18)
369
The BM biomarker present in Eq. 18 allows the analysis of the bone mass in re-
370
lation to a reference value stipulated by the user. In this equation, 1 was adopted
371
as a reference, but other values can be adopted depending on the clinical con-
372
text. Note that, in the absence of treatment, the biomarker for BMD returns to
373
the baseline value 1. The inclusion of Eq. 18 in the model makes it more flexible, 23
374
enabling site-specific BMD changes, according to the clinical data (Ross et al.,
375
2017).
376
The model parameters are summarized in Table 2. A more detailed explana-
377
tion of each parameter is found in Lemaire et al. (2004); Scheiner et al. (2013);
378
Pivonka et al. (2008).
379
2.3. Microscale strain energy density
380
The strain energy density evaluated at the microscale (microSED) is adopted
381
as a mechanoregulatory stimulus for osteocytes to drive the bone remodeling
382
process (Jacobs, 1994; R¨uberg, 2004). According to the magnitude of this energy,
383
the algorithm distinguishes remodeling zones with different behaviors.
384
The microscopic evaluation of strain energy densities follows the proposal
385
of Scheiner et al. (2013), which considers the inhomogeneous constitution of
386
the bone and the resulting inhomogeneity of its mechanical properties (Zaoui,
387
2002). The transition between macroscale and microscale involves techniques
388
of homogenization that originates from the continuum micromechanics. The
389
bone presents a hierarchical structure quite complex relative to its microscopic
390
arrangements and mechanical properties. These procedures represent a viable al-
391
ternative to obtaining effective mechanical properties of the material to be used
392
in the structural analysis. In this direction, one can cite the work of Fritsch and
393
Hellmich (2007) that proposes a multiscale material model for the bone consid-
394
ering its hierarchical structure in a multi-step homogenization procedure, re-
395
sulting in the estimation of the elastic properties of the bone at the nano and
396
sub-microscale.
397
For the sake of simplicity, we chose a characteristic length for representative
398
volume element (RVE) that allows the bone to be reasonably represented as a 24
399
biphasic composite of extravascular bone matrix with vascular pore space in be-
400
tween, as shown in Fig. 6. Each constituent has a corresponding volume fraction
401
fi , defined as fi =
402
the total volume, obeying the relation:
Vi , Vtotal
with Vi as the volume of the constituent i and Vtotal as
fbm + fvas = 1
(19)
403
with fbm as the extravascular bone matrix and fvas as vascular pores space. Due
404
to the inhomogeneities, the volume fractions of each of these phases vary accord-
405
ing to the type and bone site. In the case of cortical bone, the volume fraction
406
of the extravascular bone matrix lies within the range 0.7 - 0.95, whereas for the
407
trabecular bone the values are between 0.1 and 0.5 (Scheiner et al., 2016; Cooper
408
et al., 2007; Padilla et al., 2008; Boutroy et al., 2005, 2011).
409
410
It is possible to express fvas (or fbm ) in terms of the apparent density ρ = mtotal , Vtotal
with mtotal as total mass. Thus:
fvas = 1 −
Vbm (mtotal /ρ)
(20)
411
Assigning zero mass to the vascular pore space, i.e., mtotal = mbm , the expression
412
is reduced to: fvas = 1 −
ρ ρ or fbm = ρt ρt
(21)
413
where ρt = mbm /Vbm is the density of an imaginary bone specimen without
414
vascular pores (R¨uberg, 2004).
415
The temporal evolution of the bone constituent volume fraction is incorpo-
416
rated into the bone cells populations model. Thus, the volume fraction of ex-
417
travascular bone matrix decreases due to bone resorption and increases due to 25
418
the bone formation. In mathematical terms, this can be represented by the equa-
419
tion: df bm = kform OBa − kres OCa dt
(22)
420
with kform as the formation rate (bone mass formed per unit of time) and kres as
421
the resorption rate (bone mass resorbed per unit of time). Conversely, the volume
422
fraction of vascular pore space decreases due to bone resorption and increases
423
due the bone formation, which results in the condition: dfvas dfbm =− = −kform OBa + kres OCa dt dt
(23)
424
Both formation and resorption rates, kform and kres respectively, are consid-
425
ered species-specific cellular properties, and although known to be dependent on
426
health conditions, here they are considered invariant in time. kform is calibrated
427
so that the previously defined bone resorption rate and cells populations values
428
for stationary states imply in a balanced bone turnover (in the EDO system, we
429
have dOBp /dt = dOBa /dt = dOCa /dt = 0) thus considering the bone matrix
430
volume fraction fbm constant in Eq. 23, kf orm = kres OCa /OBa .
431
432
From the perspective of the continuum micromechanics, the homogenized fourth order stiffness of the bone Chom bone , is: Chom bone =
X
fr Cr : Aest r
(24)
r 433
where Cr is the microscopic fourth-order stiffness tensor of the constituent r,
434
with r = vas, bm, and Aest r is the estimation of the corresponding fourth-order
435
strain concentration tensor, relating the macroscopic and microscopic second or26
Figure 6: Micromechanical representation of the bone as a two-phase composite material of ”vascular” pores and ”extravascular” bone matrix with volume fractions fvas and fbm , respectively. Vascular pores host osteoblasts and osteoclasts at various differentiation stages and a large variety of biochemical factors. Artwork: Miguel T. Bahia.
436
der strain tensors. Aest r can be estimated by solving the classical matrix-inclusion
437
problem of Eshelby (Eshelby, 1957), by means of the Mori-Tanaka scheme:
Aest = r
−1 I + Pbm : r : (Cr − Cbm ) )−1 ( X −1 fs I + Pbm s : (Cr − Cbm )
(25)
s 438
where the index r denotes the phase, and the sum over the index s include both
439
phases, s = vas, bm. In addition, I is the fourth-order unit tensor with the com-
440
ponents defined by the Kronecker delta, δij = 1 for i = j and zero otherwise,
441
being Iijkl = 1/2 (δik δjl + δil δjk ), and Pbm r is the Hill tensor of phase r inserted
442
in the matrix with rigidity Cbm .
443
Volume fractions are obtained from the temporal evolution of the cellular 27
444
population model. Stiffness tensors are extracted from the work of Scheiner et al.
445
(2013).
446
A linear elastic constitutive law relates the macroscopic stress tensor acting
447
on the bone and the corresponding macroscopic strain tensor, via the macro-
448
scopic stiffness tensor (Eq. 24), i.e.: hom −1 : Σbone Σbone = Chom bone : Ebone ⇐⇒ Ebone = (Cbone )
(26)
449
The strain tensor of the extravascular bone matrix, corresponding to the
450
macroscopic loading Σbone , can be expressed with the stress concentration tensor
451
Aest bm defined in Eq. 25, that is: est hom −1 εbm = Aest : Ebone ] bm : Ebone = Abm : [(Cbone )
452
453
(27)
The microscopic strain energy density in the extravascular bone matrix is expressed by: 1 Ψbm = εbm : Cbm : εbm 2
(28)
454
The microscopic SED Ψbm obtained by the homogenization technique may
455
now be used as a mechanoregulatory function providing a representative scalar
456
value at the microscale associated with the three-dimensional macroscopic stress
457
state.
458
The bone cells populations model was implemented incorporating a FE frame-
459
work for the three-dimensional case. Each interaction of the structural analysis
460
calculates the local micro strain energy density for each node of the discretized
461
mesh, thus solving the system of differential equations that governs the tempo-
462
ral evolution of the bone cells populations. A new population of bone cells is 28
463
calculated at each node and the volume fractions are updated at each iteration.
464
The temporal evolution of the bone matrix volume fraction (Eq. 22) provides the
465
temporal evolution of apparent bone density ρ from Eq. 21. The homogenized
466
stiffness tensor Chom bone is updated for each node of the finite element mesh using
467
the new volume fractions. Note that in the first iteration, a phenomenological
468
law (Eqs. 29-30) (described in Subsection 2.4) relating the elasticity modulus and
469
the ash densities from the CT images is used in the structural analysis. The next
470
iterations use the updated homogenized stiffness tensor in the structural analy-
471
sis.
472
Bone is largely complex, both physiological and morphologically. Depending
473
on the observation scale (from macroscopic to nanoscopic), different constituents
474
with distinct mechanical properties are observed: hydroxyapatite crystals, col-
475
lagenous proteins, water, among others. Thus, to obtain the effective properties
476
of the composite (particularly the stiffness matrix) the Mori-Tanaka homogeniza-
477
tion technique was used. The actual modulus of elasticity of the bone does not
478
depend solely on ash density as the phenomenological relationship may suggest,
479
but it was adopted here for the sake of simplicity. Literature offers several ex-
480
perimental relationships between density and elastic modulus for different bones
481
(tibia, femur, spine) and different regions of the bone. Some examples are Carter
482
and Hayes (1977); Morgan et al. (2003); Snyder and Schneider (1991); Kaneko et al.
483
(2004); Keller (1994). A review of these relationships can be found in Helgason
484
et al. (2008).
485
2.4. Finite Elements Model
486
In order to apply the proposed procedure, a FE model was built from CT
487
images of a femur. The initial densities were derived from the CT images and 29
488
inputed to the FE model, as shown in Fig. 17a. We adopted a linear relationship
489
between pixel values [0.255] and bone densities.
490
There are several relationships between Young’s modulus and apparent bone
491
densities in the literature (Carter and Hayes, 1977; Morgan et al., 2003; Snyder
492
and Schneider, 1991; Kaneko et al., 2004; Keller, 1994). A review can be found in
493
Helgason et al. (2008).
494
As observed previously, in the first iteration the elasticity modulus is given
495
as a function of the ash densities derived from CT images using the relation
496
proposed by Keller (1994) and validated by Yosibash et al. (2007) and Trabelsi
497
et al. (2009):
498
Ecort = 10200ρ2.01 ash [M P a]
(29)
Etrab = 5307ρash + 469 [M P a]
(30)
Trabecular and cortical bone types are distinguished with the condition ρash >
499
0.6g/cm3 for cortical bone and ρash ≤ 0.6g/cm3 for trabecular bone (Trabelsi
500
et al., 2014).
501
The model was fully implemented in Matlab and post-processed in Paraview
502
5.0.0. The procedure for reconstruction and mapping of non-homogeneous isotropic
503
material properties is described in more details in Bahia et al. (2016).
504
The boundary conditions were based on Carter et al. (1987) and represent the
505
joint reactions and abductor muscle forces for a standard gait cycle. For each
506
cycle of FE analysis, three loading cases corresponding to the maximal adduc-
507
tion limits, maximum abduction and average support of the gait cycle were an-
508
alyzed sequentially (Fig. 8). The orientations and magnitudes of the forces were 30
(a) 3D femur mesh.
(b) Mapping of densities.
Figure 7: Discretization of the reconstructed femur in a finite element mesh with 3846 nodes and 16689 tetrahedral elements (Fig. 17a) and assignment of gray values to mesh elements (Fig. 17b). Source: Prepared by the author. Table 3: Loads used in the FE model referring to a standard gait. Orientation angles to the frontal and sagittal planes (Hambli et al., 2016; Carter et al., 1987).
Load
Orientation [deg]
Orientation [deg]
Frontal
Frontal
Sagittal
Sagittal
Case 1
2317
24
6
703
28
15
Case 2
1158
−15
35
351
−8
9
Case 3
1548
56
−20
468
35
16
509
adapted from Hambli et al. (2016) and are listed in Tab. 3. The concentrated loads
510
proposed by Hambli et al. (2016) were transformed into equivalent proportional
511
distributed loads on the femoral head and greater trochanter.
512
The strain energy density (microSED) from the combined loading cases was
513
defined as the highest nodal value. In other words, considering the three finite 31
(a) Load case 1
(b) Load case 2
(c) Load case 3
Figure 8: Load cases adopted in the 3D FE analysis. Each load specified in Tab. 3 was proportionally distributed onto the femur head and greater trochanter to simulate the stress on joint during gait. Source: Prepared by the authors.
514
element structural analyses performed for each load case, we adopted the critical
515
value (maximum absolute value) of equivalent stress obtained from the three
516
simulations in each mesh node in the bone remodeling algorithm. This choice
517
produced the maximum mechanical stimulus from the combination of the three
518
loading cases. This is justified because the static analysis does not incorporate
519
the loading history and load cycles, which usually arise in real daily activities.
520
521
The pharmacodynamic (PD) model coupling pharmacokinetics and bone remodeling proposed could be summarized in Fig. 9.
32
Figure 9: Diagram for the proposed unified bone remodeling model. A pharmacokinetic model provides the plasma concentration of the drug that contributes to the evolution of bone cells populations. Source: Prepared by the author.
522
3. RESULTS
523
The proposed PBPK model for bone seeking agents was solved by the 4th or-
524
der Runge-Kutta method and implemented in Matlab. The dosing regimen cho-
525
sen was 10 mg/day intravenous infusion of antiresorptive drug over 4 hours (20
526
h intervals). In clinical practice, intravenous infusion offers an alternative ther-
527
apeutic option with less gastrointestinal disturbances than orally administered
528
bisphosphonates (Cremers et al., 2002). 33
529
Figs. 10 and 11 illustrate the time course of drug concentration in each tissue
530
or organ for the 10 mg regimen by intravenous infusion for 4 hours per day. Note
531
that each dose (10 mg) is divided by infusion time (4 h). The organs included in the
532
model are blood, liver, kidney, stomach, well-perfused tissues (rapid perfusion),
533
poorly perfused tissue (slow perfusion), and bone. Fig. 11 presents the temporal
534
evolution of the drug separately for each organ.
535
The intravenous infusion is modeled by periodically adding the drug dose,
536
mathematically expressed here by a step or Heaviside function. The partition
537
coefficient of the kidney is greater than the coefficient of the other organs, so the
538
concentration of the drug reaches higher peaks in this organ, as can be observed
539
in Fig. 11.
540
The bone compartment includes the vascular bone surface and the extravas-
541
cular bone matrix for both cortical and trabecular bone. For this regimen, the
542
drug is not completely eliminated from the bone. In this case, the drug con-
543
centration is divided in more compartments as presented in Sec. 2.1.1. In the
544
extravascular bone matrix, a portion of the drug is not completely eliminated
545
before the next infusion, and accumulates in the tissue.
546
Fig. 12 shows the drug distribution in the four bone compartments: the cor-
547
tical and trabecular bone and their respective subdivisions, the vascular bone
548
surface and the extravascular bone matrix. The model makes it possible to cali-
549
brate the drug distribution by modifying parameters such as bone partition co-
550
efficient, drug bioavailability and rates of bone formation and resorption, given
551
by coefficients kform and kres , related to the process of bone remodeling.
552
The physiological parameters used in the simulation are listed in Tab. 1. The
553
drug concentration in the bone is given as micrograms of drug per gram of bone 34
Figure 10: Implementation of the proposed PBPK model. A dosage regimen of 10 mg of drug / day by intravenous infusion over a period of 4 hours. The chart shows the rate of infusion and amounts of drugs in each organ / tissue. Source: Prepared by the author.
554
ash and is defined as cdrug = 0.01 mcg drug/g bone ash. This value corresponds to
555
the steady state drug concentration obtained in the trabecular bone matrix with
556
the PBPK model for the 10 mg daily drug regimen intravenously for 4 h. Note
557
that the most of the physiological parameters of the PBPK model used here were
558
extracted from the literature. However, due to the lack of specific experimental
559
data, some parameters, such as bone formation and resorption rates and bone
560
partition coefficients, needed to be arbitrated by the author. It is worth noting
35
Figure 11: Concentration of drug in each organ / tissue for a dosage regimen of 10 mg of drug per day, intravenously for 4 hours.
561
that the value adopted for the drug concentration in the bone has the purpose of
562
illustrating the procedure, but needs to be validated experimentally.
563
The bone cells population model represented by Eqs. 8-18 is ilustrated on Figs.
564
13 and 14. The model was solved by numerical integration 4th order Runge-Kutta
565
method. In the first simulation (Fig. 13), the strain energy density was defined
566
as 0.13 Pa and there was no bisphosphonate stimulation. Around the 180th day
567
of iteration the population of preosteoblasts, active osteoblasts and osteoclasts
568
reached the steady state condition. At the end of 200 days the BMD biomarker 36
Figure 12: Drug distribution in the trabecular and cortical bones for a dosage regimen of 10 mg of drug /day intravenously for 4 hours.
569
had an increase of 1.98%.
570
Note that cell populations maintain the steady state value when disturbance
571
is removed from the system of linear differential equations. To analyze the im-
572
pact of the system perturbations such as the inclusion of RANKL or OPG, it is
573
more significant to evaluate the system’s behavior in terms of percent change
574
from the current state relative to the initial reference values. After the stimu-
575
lus, as the pharmacological and/or mechanical, a new population configuration
576
is obtained, and the disturbed system steady state value will be different from
577
the initial steady state value, as shown in Fig. 13. Removing the system dis-
578
turbance, the cells populations return to the initial values, as shown by Lemaire
579
et al. (2004).
580
The second simulation aims to assess the impact of the anti-resortive agent 37
Figure 13: Dynamics of bone cell populations formed by pre-osteoblasts, active osteoblasts and active osteoclasts, and bone mass density (BMD) evolution. Mechanical stimulation by SED. Without anti-resorbtive agent.
581
on bone cells population and BMD biomarker, as shown in Fig. 14. The me-
582
chanical stimulus was maintained. The antiresorptive effect was simulated as
583
an increase in the rate of apoptosis of active osteoclast and the inclusion of a
584
regulatory function in RANKL expression. We observed that the increase of os-
585
teoclast apoptosis rate was not sufficient to increase the rate of differentiation
586
of precursor osteoblasts in the model, so we added another term (a regulatory
587
function given by Eq. 16) to the RANKL expression to produce the desired effect.
588
Repeating the simulation with the pharmacological stimulus after the 20th day,
589
we observed an increase of 3.28% in BMD biomarkers, slightly higher than that
38
590
observed in the previous simulation. The sensitivity of the BMD biomarker to the
591
mechanical stimulus, that changes the rate of differentiation of pre osteoblasts
592
was higher than its sensitivity to the pharmacological stimulus, that acts on the
593
apoptosis of osteoclasts and the expression of RANKL in the model. This suggests
594
that other pathways could be considered in the expression of pharmacological
595
intervention, including adjustments in mechanoregulation function. It is reason-
596
able to assume changes in sensitivity to mechanical stimulus in the presence of
597
certain bone diseases such as osteoporosis.
Figure 14: Dynamics of bone cell populations formed by pre-osteoblasts, active osteoblasts and active osteoclasts, and bone mass density (BMD) evolution. Mechanical stimulation by SED. With anti-resorptive agent.
598
The expression that defines the differentiation of pre-osteoblasts is a function 39
599
of the concentration of active osteoclasts. Thus, the increase in active osteoclasts
600
apoptosis does not increase the population of active osteoblasts, but reduces
601
the rate of differentiation of precursor osteoblasts. The increase in osteoclast-
602
osteoblast ratio was observed when the drug-regulatory function (Eq. 16) was
603
included in Eq. 13.
604
3.1. Postmenopausal osteoporosis treatment simulation
605
Postmenopausal osteoporosis alters bone microarchitecture, resulting in in-
606
creased porosity, reduced mechanical resistance and increased risk of fractures.
607
In order to verify the validity of the model in this scenario, we have simulated
608
postmenopausal osteoporosis and the effect of a drug treatment. The increased
609
porosity can be simulated in different ways in our proposed model. For instance,
610
by reducing the minimum rate of OPG production per cell, or modifying the rate
611
of administration of PTH and/or RANKL. For simplicity, here we chose to reduce
612
the minimum rate of OPG production per cell, according to Lemaire et al. (2004).
613
Considering that the bone adaptive response to mechanical stimuli is drasti-
614
cally affected by the changes in the RANK-RANKL-OPG system caused by osteo-
615
porosis, the anabolic factor λ in the mechanoregulation equation was calibrated
616
to favor osteoclast activity while decreasing osteoblast activity, which results in
617
greater bone resorption. Changes in anabolic factor follow the work of Scheiner
618
et al. (2013).
619
Bisphosphonates acts on osteoporosis by preventing the fixation of osteo-
620
clasts in the bone matrix and increasing osteoclastic apoptosis, thus inhibiting
621
bone resoption.
622
Fig. 15 illustrates the changes in the bone matrix volume fraction over time,
623
with and without antiresorptive agents. Pharmacological treatment was initiated 40
Figure 15: Temporal evolution of the bone matrix volume fraction with and without antiresorptive treatment. Pharmacological treatment started after the 200◦ day. Initial volume fraction fbm equal to 0.95.
624
after the 200◦ day. The minimum rate of OPG production per cell, represented by
625
the term koP , was reduced by half, i.e. 1.105 pM.day−1 /pM cells, and the anabolic
626
factor λ was set to 1.2. Pharmacological treatment was simulated by assigning
627
non-zero values to the variable cdrug that corresponds to the anti-resorptive drug
628
concentration in the bone, Eq. 11. The temporal evolution of bone matrix volume
629
fractions fbm for cdrug = 0 (no pharmacological stimulus), cdrug = 0.01, 0.1 and
630
1.0 mg/g bone ash is given in Fig. 15. The initial value for bone matrix volume
631
fraction fbm was 0.95. The increase in the pharmacological stimulus produced
41
632
a smaller decrease of the volume fraction, which is expected for this type of
633
treatment. The simulation was investigated for a period of 1200 days.
634
When the cell concentration reaches a plateau, the volume fraction starts de-
635
caying linearly. This occurs because the rates of bone formation and resorption,
636
kres and kform , which control the temporal variation of the bone matrix volume
637
fraction were considered constant in this work.
638
3.2. Bone remodeling simulation
639
In this section, we present the implementation of the proposed procedure as
640
shown in Fig. 9. Due to practical and ethical reasons, obtaining in vivo informa-
641
tion about drug disposition in human bones is challenging. Specimen collection
642
can damage bone tissue and is rarely conducted in the clinical setting Stepensky
643
et al. (2003). Drug disposition also depends on the type of anti-resorptive agent,
644
with significant differences even among bisphosphonates, such as alendronate
645
and etyledronate Masarachia et al. (1996). For this reason, the authors adopted
646
only one reference value for the simulations, in order to qualitatively evaluate
647
the procedure. Drug concentration in the bone used as the pharmacological stim-
648
ulus was cdrug = 0.5mcg/g bone ash. This value roughly corresponds to the drug
649
concentration in the bone matrix obtained for the 10 mg daily dosing regimen
650
via 4 h intravenous infusion as shown in Fig. 11.
651
Each procedure iteration involves a structural analysis using finite element
652
method, numerical integration of the system of differential equations that gov-
653
erns the dynamics of cells populations, and update of the mechanical properties.
654
The temporal evolution of the cells populations is performed in cycles of 120
655
days. Figure 17 shows the distribution of densities in a longitudinal section of
656
the reconstructed femur for the initial and final configurations after 4 iterations 42
(a) Longitudinal section plane.
(b) Cross section plane.
Figure 16: Cutting planes adopted in the visualization. The length of femur piece is about 40 cm and the cross-section was made at a height of 10 cm measured from the lower end.
(a) Initial configuration.
(b) Final configuration after 360 cycles.
Figure 17: Density distribution in the femur at baseline and after bone remodeling procedure.
43
(a) Initial densities
(b) K-means clustering initial
(c) Densities after 120 cycles
(d) K-means clustering 120 cycles
(e) Densities after 240 cycles
(f) K-means clustering 240 cycles
Figure 18: Cross section of the femur in the diaphyseal region, 60 mm above the femur’s lower end: (a), (c) and (e) are densities distribution at the initial point, after 120 and 240 cycles, respectively; (b), (d) and (f) correspond to the image segmentation at the initial point, after 120 and 240 cycles, respectively. K-means clustering algorithm was adopted to evaluate the percent increase of the cortical region (defined here as the sum of areas 1 (yellow) and 2 (grey) in (b), (d) and (f)).
44
Table 4: Average ash density (µ), standard deviation (σ) in [g/cm3 ] and percent of each bone type over the procedure’s iterations for the transversal section of the femur illustrated in Fig. 18. Source: Elaborated by the author.
Bone type
Initial configuration
120 cycles 3
240 cycles 3
Percentage
µ ± σ[g/cm ]
Percentage
µ ± σ[g/cm ]
Percentage
µ ± σ[g/cm3 ]
Trabecular
28.11%
0.437 ± 0.098
23.32%
0.502 ± 0.066
17.24%
0.477 ± 0.098
Cortical
71.89%
1.147 ± 0.257
76.67%
1.043 ± 0.179
82.76%
1.245 ± 0.207
Figure 19: Algorithm convergence after approximately 360 iterations. Source: Prepared by the author.
657
or 480 days.
658
Fig. 18 ilustrates a cross section of the femur in the diaphyseal region af-
659
ter convergence. We observe an increase in cortical bone area of approximately
660
7.68% relative to the initial area. The increase was estimated using K-means clus-
661
tering to segment the image and count the number of pixels in each segmented
662
region (adding up areas 1 and 2 shown in Fig. 18). Convergence is shown in
663
Fig. 19 45
(b) Initial densities distribution
(a) Cutting plan
(c) Densities distribution after 240 cycles
Figure 20: Cutting plane in the neck region of the femur and views of densities distribution at the initial and final configurations.
664
Bone remodeling in the neck region of the femur obtained by numerical sim-
665
ulation is illustrated in Fig. 20. Significant change in density distribution and an
666
increase in cortical bone thickness can be observed. The new density distribution
667
depends greatly on the boundary conditions adopted in the simulation.
668
4. DISCUSSION
669
Here we show a mathematical model for bone remodeling that integrates the
670
bone cells populations dynamics and a physiologically based pharmacokinetic
671
model. This is a first attempt to incorporate recent advances in mechanobiology 46
672
concerning the molecular pathways that regulate bone remodeling into a model
673
that can contribute to the development of new antiresorptive or anabolic drugs.
674
In the future, the platform presented here can evolve to precision medicine, al-
675
lowing safer and more effective therapeutic treatments.
676
Here we adopt the Lemaire proposal, which include the RANK-RANKL-OPG
677
pathway, as the core of our model. Lemaire’s model was modified to include
678
the mechanical stimulus evaluated at the microscale and a pharmacoregulator
679
function related to the drug concentration at the bone sites.
680
More sophisticated models of the dynamics of bone cells have been published
681
including other signaling molecules. However, the inclusion of more variables in
682
the model should be pondered with caution since the calibration of parameters
683
can become very difficult or even unfeasible.
684
The changes in bone densities distribution, in the cortical bone area and the
685
increase in the BMD biomarker obtained in the simulations indicate that our in-
686
tegrated model captures reasonably well the expected adaptive behavior of bone
687
during remodeling when subjected to mechanical and pharmacological stimuli.
688
A transversal section at the diaphyseal region after convergence of the algorithm
689
showed an increase in the cortical bone area of 7.68% relative to the initial area.
690
The bone cells populations model includes an anti-resorptive agent, showed a
691
BMD biomarker increase of 3.28%, slightly higher than that obtained in the sim-
692
ulation with mechanical stimuli only. This difference was smaller than expected,
693
suggesting that the implementation is more sensitive to mechanical than phar-
694
macological stimulus. Note that the mechanical stimulus is represented in the
695
model as a mechanoregulator function added to the precursor osteoblasts ex-
696
pression stimulating their differentiation into active osteoblasts. Thus, the study 47
697
indicates that the pharmacological stimuli should be better represented by an
698
increase in the rate of proliferation of precursor osteoblasts, controlled, for in-
699
stance, by the calibration of λ in the mechanoregulator function. Indeed, in a sce-
700
nario of osteoporosis, a decreased sensitivity to mechanical loading and damage
701
is associated to the decline in osteocyte density (Geissler et al., 2015; Frost, 1964).
702
Note that osteocytes and their processes make up a highly sensitive mechanosen-
703
sory system (Cowin et al., 1991) that affects the differentiation of precursors os-
704
teoblasts and consequently the bone formation by active osteoblasts. Future im-
705
plementations could explore other ways to stimulate cell differentiation.
706
An increase in the rate of osteoclasts apoptosis produces a decrease in the rate
707
of differentiation of pre-osteoblasts. Therefore, the anti-resorptive action should
708
not be based solely on the increase in the rate of osteoclastic apoptosis as this is
709
not sufficient to represent an increase in BMD obtained with drug administration.
710
This effect was obtained with a modification in RANKL expression including a
711
regulatory function corresponding to the presence of the drug.
712
According to the parameters adopted in the simulation, the pharmacological
713
stimulus (pharmaco-regulatory function) are normalized between [0,1], regard-
714
less of drug concentration. In real situations, there is a therapeutic range related
715
to the mechanism of action of each drug, in which a maximum response with
716
minimal toxicity is obtained for most patients (Rosenbaum, 2016).
717
Further research should focus on the calibration of the model parameters to
718
better reproduce the behavior of commercially available antiresorptive medica-
719
tions for the treatment of osteoporosis such as alendronate, and risedronate.
720
In the proposed model the choice of boundary conditions strongly influences
721
the cortical formation in the process of bone remodeling due to the mechanical 48
722
stimulus via microscopic strain energy density. Because we adopted a mechan-
723
ical loading less complex than the physiological loading, the cortical thickness
724
did not increase in a strictly concentric manner and presented certain variations,
725
as can be observed in Fig. 18. Thus, it is more plausible to evaluate the increase
726
of cortical area by means of image segmentation than to measure its thickness
727
since the latter varies largely depending on the measurement angle adopted. The
728
increase of cortical bone area was measured through image segmentation using
729
K-means clustering, which allowed a practical quantitative assessment of the
730
procedure’s effectiveness.
731
The mechanoregulation function used the microscopic strain energy density
732
instead of the macroscopic one. As Scheiner et al. (2013) points out, there is a sig-
733
nificant dependence of macroscopic stress state on the current vascular porosity
734
fvas . In the case of close-to-zero porosity, the two energy densities (at the macro
735
and microscale) tend to coincide, while for high porosities, large deviations are
736
found, and the value of macroscopic strain energy density (macroSED) underesti-
737
mates the value of microSED. Also, adopting a stimulus at the microscale makes
738
more sense in this model because inhibition or activation of bone remodeling
739
occur at the cellular level.
740
One limitation of the model concerns the representation of the bone as a
741
nonhomogeneous isotropic material. Although bones present an anisotropic be-
742
havior, the model showed good agreement with the bone’s expected adaptive
743
behavior and captured reasonably well both the cells interactions and molecu-
744
lar pathways that regulate the bone remodeling. In the case of subject-specific
745
FE models, the adoption of a nonhomogeneous isotropic distribution of material
746
properties derived from CT data, as performed here is preferable to average val49
747
ues reported in the literature. This choice leads to more accurate results, since the
748
distribution of stresses in bone is highly related to the distribution of mechanical
749
properties in the bone tissue (Taddei et al., 2004).
750
The PBPK model produced different distributions of drug concentration for
751
each bone type, resembling a real situation. Modeling the long term effects of
752
bone seeking agents as bisphosphonates remains a challenge. The complete elim-
753
ination of bisphosphonate from the body, for instance, is reported to take 10 years
754
on average. Note that the time scale for pharmacological models is defined, in
755
general, in hours, while bone remodeling occurs in days, months and years.
756
As previously mentioned, the PBPK model allows the assessment of differ-
757
ent drug administration regimens in order to optimize treatments, ensure ther-
758
apeutic drug dosages, and minimize possible side effects and liver overload. For
759
instance, oral intake requires a much larger amount of drug to achieve the same
760
concentration obtained with a constant intravenous infusion, which might re-
761
sult in more side effects due metabolisation in the liver. The pharmacological
762
model typically deals with time scales much smaller than the remodeling model
763
time scale (months and years); for this reason we use only the local steady state
764
drug concentration in the expression of pharmacological stimulus. However,
765
it should be noted that local steady-state drug concentration depends on fac-
766
tors such as administration regimen (route, dose frequency), drug characteris-
767
tics (bioavailability, half-life, form of action), and individual patient characteris-
768
tics (weight, height, and health condition). Such elements are not directly con-
769
templated by the remodeling model, but they may be partially contemplated in
770
the physiologically-based pharmacokinetic model. As seen in Section 2.1.1, the
771
proposed pharmacological model provides the temporal evolution of drug con50
772
centration in 4 distinct sub-compartments: cortical and trabecular extravascular
773
bone matrix, and cortical and trabecular vascularized bone surface. A uniform
774
distribution of the drug in each compartment is an inherent assumption in com-
775
partmental models. However, given the extreme morphological and physiologi-
776
cal complexity of the bone, the actual spatial distribution of the drug across the
777
various bones that make up the human skeleton is not uniform. Nonetheless, the
778
authors understand that the fact that local drug concentration can be discrimi-
779
nated in 4 different regions of bone tissue represents a significant advance over
780
current remodeling models, which generally lack this information. Future efforts
781
may move toward exploring local aspects of drug delivery.
782
Results of the bone remodeling simulation agree qualitatively with the ex-
783
perimental results that show an increase in bone mass due to physical exercise
784
and pharmacological treatment. The beneficial effects on BMD associated with
785
exercise vary according to exercise modality (impact and aerobic, resistance, bal-
786
ance and proprioception, whole body vibration, aquatic) and intensity, which
787
renders a quantitative comparison challenging. Ma et al. (2013) found significant
788
and positive effects on the BMD of the femur neck in postmenopausal and per-
789
imenopausal women with walking as a single exercise therapy. Another study
790
(Velez et al., 2008) comparing elderly runners with a sedentary control group
791
found a significantly better whole body BMD for the runner group. Also, a
792
high-intensity resistance and impact training program enhances indices of bone
793
strength and functional performance in postmenopausal women with low bone
794
mass, according to (Watson et al., 2018). Regarding pharmacological treatment,
795
Chailurkit et al. (2003) concluded that treatment with alendronate at a regimen of
796
10 mg daily for postmenopausal women with osteoporosis significantly increased 51
797
BMD at all skeletal sites. According to Liberman et al. (1995), the difference in
798
BMD between women receiving this dosage and those receiving placebo was 5.9
799
0.5 percent in the femoral neck and 7.8 0.6 percent in the trochanter. Fig. 20
800
shows an axial section of the femoral neck and the cortical thickening obtained
801
in the remodeling simulation qualitatively agreeing with the experimental results
802
mentioned above.
803
Here we adopted the steady-state value of drug concentration produced by
804
the PBPK model as an input for the bone remodeling algorithm. Notice also
805
that the concentration of the drug in the bone adopted here is intended only
806
to illustrate an application of the model. Further studies are needed in order to
807
compare our results with real clinical data.
808
5. CONCLUSION
809
Here, we describe a mathematical model for bone remodeling integrating dy-
810
namics of bone cells populations and a physiologically based pharmacokinetics
811
model. The results of the simulations show an increase in cortical bone area,
812
considering a transversal section in the diaphyseal region and an increase in the
813
BMD biomarker evaluated by the system of ordinary differential equations. The
814
model seems to be more sensitive to the mechanical stimulus than to the phar-
815
macological one, suggesting that other signaling paths can be explored in future
816
implementations. Here, the pharmacological stimulus was given by the modifica-
817
tion of the RANKL expression and the increase in osteoclast apoptosis. The study
818
suggests that the differentiation of precursor osteoblasts must also be modified
819
to better reproduce the drug effect (for instance, by modifying the mechanoreg-
820
ulator expression). In conclusion, the model is potentially effective in predicting 52
821
the effects of bone-seeking agents used in bone disease treatments, like alen-
822
dronate, and ziledronate. Further research should focus on the calibration of
823
model parameters to better reproduce the behavior of commercially available
824
antiresorptive medications.
825
Acknowledgements
826
The authors would like to thank Prof. Monica Campiteli of Instituto de F´ısica
827
de S˜ao Carlos of the Universidade de S˜ao Paulo for suggestions and revision of
828
the manuscript.
829
Disclosure statement
830
No conflict of interest is declared.
831
Funding
832
None
833
834
Ethical approval Not required
835
REFERENCES
836
Ara´ujo, D. V., de Oliveira, J. H., Bracco, O. L., 2005. Cost of osteoporotic hip
837
fracture in the brazilian private health care system. Arquivos Brasileiros de
838
Endocrinologia & Metabologia 49 (6), 897–901.
53
839
Bahia, M. T., Hecke, M. B., Mercuri, E. G., 2019. Image-based anatomical recon-
840
struction and pharmaco-mediated bone remodeling model applied to a femur
841
with subtrochanteric fracture: A subject-specific finite element study. Medical
842
Engineering & Physics (69), 58–71.
843
Bahia, M. T., Mercuri, E. G., Hecke, M. B., 2016. Fe analyses of a reconstructed
844
femur with mapping of inhomogeneous isotropic material properties from ct
845
images. VII European Congress on Computational Methods in Applied Sci-
846
ences and Engineering, Crete Island, Greece.
847
Barkaoui, A., 2012. Mod´elisation multi´echelle du comportement m´ecano-
848
biologique de l’os humain: de l’ultrastructure au remodelage osseux. Ph.D.
849
thesis, Universit´e d’Orl´eans.
850
Boutroy, S., Bouxsein, M. L., Munoz, F., Delmas, P. D., 2005. In vivo assessment of
851
trabecular bone microarchitecture by high-resolution peripheral quantitative
852
computed tomography. The Journal of Clinical Endocrinology & Metabolism
853
90 (12), 6508–6515.
854
Boutroy, S., Vilayphiou, N., Roux, J.-P., Delmas, P. D., Blain, H., Chapurlat, R. D.,
855
Chavassieux, P., 2011. Comparison of 2d and 3d bone microarchitecture eval-
856
uation at the femoral neck, among postmenopausal women with hip fracture
857
or hip osteoarthritis. Bone 49 (5), 1055–1061.
858
Brown, R. P., Delp, M. D., Lindstedt, S. L., Rhomberg, L. R., Beliles, R. P.,
859
1997. Physiological parameter values for physiologically based pharmacoki-
860
netic models. Toxicology and industrial health 13 (4), 407–484.
861
Burr, D. B., Allen, M. R., 2013. Basic and applied bone biology. Academic Press. 54
862
Carter, D., Fyhrie, D. P., Whalen, R., 1987. Trabecular bone density and loading
863
history: regulation of connective tissue biology by mechanical energy. Journal
864
of Biomechanics 20 (8), 785789–787794.
865
866
Carter, D. R., Hayes, W. C., 1977. The compressive behavior of bone as a twophase porous structure. The Journal of Bone & Joint Surgery 59 (7), 954–962.
867
Chailurkit, L.-o., Jongjaroenprasert, W., Rungbunnapun, S., Ongphiphadhanakul,
868
B., Sae-Tung, S., Rajatanavin, R., 2003. Effect of alendronate on bone mineral
869
density and bone turnover in thai postmenopausal osteoporosis. Journal of
870
bone and mineral metabolism 21 (6), 421–427.
871
Cooper, D. M., Thomas, C. D. L., Clement, J. G., Turinsky, A. L., Sensen, C. W.,
872
Hallgr´ımsson, B., 2007. Age-dependent change in the 3d structure of cortical
873
porosity at the human femoral midshaft. Bone 40 (4), 957–965.
874
Cowin, S. C., Moss-Salentijn, L., Moss, M. L., 1991. Candidates for the
875
mechanosensory system in bone. Journal of biomechanical engineering
876
113 (2), 191–197.
877
Cremers, S., Sparidans, R., den Hartigh, J., Hamdy, N., Vermeij, P., Papapoulos,
878
S., 2002. A pharmacokinetic and pharmacodynamic model for intravenous bis-
879
phosphonate (pamidronate) in osteoporosis. European journal of clinical phar-
880
macology 57 (12), 883–890.
881
882
Doblar´e, M., Garcıa, J., 2002. Anisotropic bone remodelling model based on a continuum damage-repair theory. Journal of Biomechanics 35 (1), 1–17.
883
Eshelby, J. D., 1957. The determination of the elastic field of an ellipsoidal inclu-
884
sion, and related problems. In: Proceedings of the Royal Society of London A: 55
885
Mathematical, Physical and Engineering Sciences. Vol. 241. The Royal Society,
886
pp. 376–396.
887
´ M., Raffaelli, M. P., Bracco, O. L., Takata, E. T., Reis, F. B., Santili, C., Fortes, E.
888
Lazaretti-Castro, M., 2008. High morbid-mortability and reduced level of os-
889
teoporosis diagnosis among elderly people who had hip fractures in s˜ao paulo
890
city. Arquivos Brasileiros de Endocrinologia & Metabologia 52 (7), 1106–1114.
891
Fritsch, A., Hellmich, C., 2007. fiuniversalfimicrostructural patterns in
892
cortical and trabecular, extracellular and extravascular bone materials:
893
Micromechanics-based prediction of anisotropic elasticity. Journal of Theoret-
894
ical Biology 244 (4), 597–620.
895
Frost, H. M., 1964. Dynamics of bone remodeling. Bone biodynamics 315.
896
Ganghoffer, J.-F., 2012. A contribution to the mechanics and thermodynamics of
897
surface growth. application to bone external remodeling. International Journal
898
of Engineering Science 50 (1), 166–191.
899
Garc´ıa-Aznar, J., Rueberg, T., Doblare, M., 2005. A bone remodelling model cou-
900
pling microdamage growth and repair by 3D BMU-activity. Biomechanics and
901
modeling in mechanobiology 4 (2-3), 147–167.
902
Geissler, J. R., Bajaj, D., Fritton, J. C., 2015. American society of biomechanics
903
journal of biomechanics award 2013: cortical bone tissue mechanical quality
904
and biological mechanisms possibly underlying atypical fractures. Journal of
905
Biomechanics 48 (6), 883–894.
906
907
Gieschke, R., Serafin, D., 2013. Development of innovative drugs via modeling with MATLAB. Springer. 56
908
Goda, I., Ganghoffer, J.-F., Maurice, G., 2016. Combined bone internal and ex-
909
ternal remodeling based on Eshelby stress. International Journal of Solids and
910
Structures 94, 138–157.
911
Gowder, S., 2012. Cell Interaction. IntechOpen.
912
URL https://books.google.es/books?id=8hOeDwAAQBAJ
913
Hambli, R., Boughattas, M. H., Daniel, J.-L., Kourta, A., 2016. Prediction of deno-
914
sumab effects on bone remodeling: A combined pharmacokinetics and finite
915
element modeling. Journal of the mechanical behavior of biomedical materials
916
60, 492–504.
917
Helgason, B., Perilli, E., Schileo, E., Taddei, F., Brynj´olfsson, S., Viceconti, M.,
918
2008. Mathematical relationships between bone density and mechanical prop-
919
erties: a literature review. Clinical Biomechanics 23 (2), 135–146.
920
Hofbauer, L., Kuhne, C., Viereck, V., et al., 2004. The opg/rankl/rank system in
921
metabolic bone diseases. Journal of Musculoskeletal and Neuronal Interactions
922
4 (3), 268.
923
ICRP, 1975. Report of the Task Group on Reference Man. ICRP Publication 23.
924
Pergamon Press, Oxford, international Commission on Radiological Protec-
925
tion.
926
927
Jacobs, C. R., 1994. Numerical simulation of bone adaptation to mechanical loading. Ph.D. thesis, Stanford University, Stanford, EUA.
928
Jones, D., Nolte, H., Schol¨ubbers, J., Turner, E., Veltel, D., 1991. Biochemical signal
929
transduction of mechanical strain in osteoblast-like cells. Biomaterials 12 (2),
930
101–110. 57
931
Kaneko, T. S., Bell, J. S., Pejcic, M. R., Tehranzadeh, J., Keyak, J. H., 2004. Mechan-
932
ical properties, density and quantitative ct scan data of trabecular bone with
933
and without metastases. Journal of Biomechanics 37 (4), 523–530.
934
Kaspar, D., Seidl, W., Neidlinger-Wilke, C., Beck, A., Claes, L., Ignatius, A., 2002.
935
Proliferation of human-derived osteoblast-like cells depends on the cycle num-
936
ber and frequency of uniaxial strain. Journal of Biomechanics 35 (7), 873–880.
937
Keller, T. S., 1994. Predicting the compressive mechanical behavior of bone. Jour-
938
nal of Biomechanics 27 (9), 1159–1168.
939
Komarova, S. V., Smith, R. J., Dixon, S. J., Sims, S. M., Wahl, L. M., 2003. Mathe-
940
matical model predicts a critical role for osteoclast autocrine regulation in the
941
control of bone remodeling. Bone 33 (2), 206–215.
942
943
Kroll, M. H., 2000. Parathyroid hormone temporal effects on bone formation and resorption. Bulletin of mathematical biology 62 (1), 163–188.
944
Lemaire, V., Tobin, F. L., Greller, L. D., Cho, C. R., Suva, L. J., 2004. Modeling the
945
interactions between osteoblast and osteoclast activities in bone remodeling.
946
Journal of theoretical biology 229 (3), 293–309.
947
Liberman, U. A., Weiss, S. R., Br¨oll, J., Minne, H. W., Quan, H., Bell, N. H.,
948
Rodriguez-Portales, J., Downs Jr, R. W., Dequeker, J., Favus, M., et al., 1995. Ef-
949
fect of oral alendronate on bone mineral density and the incidence of fractures
950
in postmenopausal osteoporosis. New England Journal of Medicine 333 (22),
951
1437–1444.
952
Lin, J., Duggan, D., Chen, I.-W., Ellsworth, R., 1991. Physiological disposition of 58
953
alendronate, a potent anti-osteolytic bisphosphonate, in laboratory animals.
954
Drug Metabolism and Disposition 19 (5), 926–932.
955
Lin, J. H., Chen, I., Deluna, F. A., Hichens, M., 1992. Renal handling of alen-
956
dronate in rats. an uncharacterized renal transport system. Drug metabolism
957
and disposition 20 (4), 608–613.
958
Louna, Z., Goda, I., Ganghoffer, J.-F., 2019. Homogenized strain gradient remod-
959
eling model for trabecular bone microstructures. Continuum Mechanics and
960
Thermodynamics, 1–29.
961
Ma, D., Wu, L., He, Z., 2013. Effects of walking on the preservation of bone min-
962
eral density in perimenopausal and postmenopausal women: a systematic re-
963
view and meta-analysis. Menopause 20 (11), 1216–1226.
964
Martin, T., 2004. Paracrine regulation of osteoclast formation and activity: mile-
965
stones in discovery. Journal of Musculoskeletal and Neuronal Interactions 4 (3),
966
243.
967
Masarachia, P., Weinreb, M., Balena, R., Rodan, G., 1996. Comparison of the dis-
968
tribution of 3h-alendronate and 3h-etidronate in rat and mouse bones. Bone
969
19 (3), 281–290.
970
McMullin, T. S., 2005. Drugs. In: Reddy, M., Yang, R., Andersen, M. E., Clewell
971
III III, H. J. (Eds.), Physiologically based pharmacokinetic modeling: science
972
and applications. John Wiley & Sons, New Jersey, Ch. 10, pp. 273–296.
973
974
Miller, P. D., 2009. Denosumab: anti-rankl antibody. Current osteoporosis reports 7 (1), 18–22. 59
975
Morgan, E. F., Bayraktar, H. H., Keaveny, T. M., 2003. Trabecular bone modulus–
976
density relationships depend on anatomic site. Journal of Biomechanics 36 (7),
977
897–904.
978
979
980
981
Mori, T., Tanaka, K., 1973. Average stress in matrix and average elastic energy of materials with misfitting inclusions. Acta metallurgica 21 (5), 571–574. Nestorov, I., 2003. Whole body pharmacokinetic models. Clinical pharmacokinetics 42 (10), 883–908.
982
Nestorov, I. A., Aarons, L. J., Arundel, P. A., Rowland, M., 1998. Lumping of
983
whole-body physiologically based pharmacokinetic models. Journal of phar-
984
macokinetics and biopharmaceutics 26 (1), 21–46.
985
NIH, 2001. Nih consensus development panel on osteoporosis prevention, diag-
986
nosis, and therapy, march 7-29, 2000: highlights of the conference. South Med
987
J 94 (6), 569–573.
988
O’Flaherty, E. J., 1991a. Physiologically based models for bone-seeking elements:
989
I. rat skeletal and bone growth. Toxicology and applied pharmacology 111 (2),
990
299–312.
991
O’Flaherty, E. J., 1991b. Physiologically based models for bone-seeking elements:
992
Iii. human skeletal and bone growth. Toxicology and applied pharmacology
993
111 (2), 332–341.
994
Padilla, F., Jenson, F., Bousson, V., Peyrin, F., Laugier, P., 2008. Relationships of
995
trabecular bone structure with quantitative ultrasound parameters: In vitro
996
study on human proximal femur using transmission and backscatter measure-
997
ments. Bone 42 (6), 1193–1202. 60
998
Pastrama, M.-I., Scheiner, S., Pivonka, P., Hellmich, C., 2018. A mathemati-
999
cal multiscale model of bone remodeling, accounting for pore space-specific
1000
mechanosensation. Bone 107, 208–221.
1001
Pertinez, H., Chenel, M., Aarons, L., 2013. A physiologically based pharmacoki-
1002
netic model for strontium exposure in rat. Pharmaceutical research 30 (6),
1003
1536–1552.
1004
1005
Pinheiro, M. d. M., 2008. Mortalidade ap´os fratura por osteoporose. Arquivos Brasileiros de Endocrinologia & Metabologia 52 (7), 1071–1072.
1006
Pinheiro, M. d. M., Eis, S. R., 2010. Epidemiology of osteoporotic fractures in
1007
brazil: what we have and what we need. Arquivos Brasileiros de Endocrinolo-
1008
gia & Metabologia 54 (2), 164–170.
1009
Pivonka, P., Zimak, J., Smith, D. W., Gardiner, B. S., Dunstan, C. R., Sims, N. A.,
1010
Martin, T. J., Mundy, G. R., 2008. Model structure and control of bone remod-
1011
eling: a theoretical study. Bone 43 (2), 249–263.
1012
1013
Porras, A. G., Holland, S. D., Gertz, B. J., 1999. Pharmacokinetics of alendronate. Clinical pharmacokinetics 36 (5), 315–328.
1014
Qin, Q.-H., Wang, Y.-N., 2012. A mathematical model of cortical bone remodel-
1015
ing at cellular level under mechanical stimulus. Acta Mechanica Sinica 28 (6),
1016
1678–1692.
1017
Rieger, R., 2011. Mod´elisation m´ecano-biologique par e´ l´ements finis de l’os
1018
trab´eculaire: des activit´es cellulaires au remodelage osseux. Ph.D. thesis, Uni-
1019
versit´e d’Orl´eans. 61
1020
1021
Rosenbaum, S. E., 2016. Basic pharmacokinetics and pharmacodynamics: An integrated textbook and computer simulations. John Wiley & Sons.
1022
Ross, D. S., Mehta, K., Cabal, A., 2017. Mathematical model of bone remodeling
1023
captures the antiresorptive and anabolic actions of various therapies. Bulletin
1024
of mathematical biology 79 (1), 117–142.
1025
1026
R¨uberg, T., 2004. Computer simulation of adaptive bone remodeling. Ph.D. thesis, Technische Universit¨at Braunschweig, Germany.
1027
Scheiner, S., Pivonka, P., Hellmich, C., 2013. Coupling systems biology with mul-
1028
tiscale mechanics, for computer simulations of bone remodeling. Computer
1029
Methods in Applied Mechanics and Engineering 254, 181–196.
1030
Scheiner, S., Pivonka, P., Hellmich, C., 2016. Poromicromechanics reveals that
1031
physiological bone strains induce osteocyte-stimulating lacunar pressure.
1032
Biomechanics and modeling in mechanobiology 15 (1), 9–28.
1033
Scheiner, S., Pivonka, P., Smith, D., Dunstan, C., Hellmich, C., 2014. Mathe-
1034
matical modeling of postmenopausal osteoporosis and its treatment by the
1035
anti-catabolic drug denosumab. International journal for numerical methods
1036
in biomedical engineering 30 (1), 1–27.
1037
Snyder, S. M., Schneider, E., 1991. Estimation of mechanical properties of cortical
1038
bone by computed tomography. Journal of Orthopaedic Research 9 (3), 422–
1039
431.
1040
1041
Stepensky, D., Kleinberg, L., Hoffman, A., 2003. Bone as an effect compartment. Clinical pharmacokinetics 42 (10), 863–881. 62
1042
Szulc, P., Bouxsein, M. L., 2011. Overview of osteoporosis: Epidemiology and
1043
clinical management. Vertebral Fracture Initiative Resource Document PART
1044
I.
1045
Taddei, F., Pancanti, A., Viceconti, M., 2004. An improved method for the auto-
1046
matic mapping of computed tomography numbers onto finite element models.
1047
Medical engineering & physics 26 (1), 61–69.
1048
Trabelsi, N., Milgrom, C., Yosibash, Z., 2014. Patient-specific fe analyses of
1049
metatarsal bones with inhomogeneous isotropic material properties. Journal
1050
of the mechanical behavior of biomedical materials 29, 177–189.
1051
Trabelsi, N., Yosibash, Z., Milgrom, C., 2009. Validation of subject-specific au-
1052
tomated p-fe analysis of the proximal femur. Journal of Biomechanics 42 (3),
1053
234–241.
1054
Velez, N., Zhang, A., Stone, B., Perera, S., Miller, M., Greenspan, S., 2008. The
1055
effect of moderate impact exercise on skeletal integrity in master athletes. Os-
1056
teoporosis International 19 (10), 1457–1464.
1057
Watson, S. L., Weeks, B. K., Weis, L. J., Harding, A. T., Horan, S. A., Beck, B. R.,
1058
2018. High-intensity resistance and impact training improves bone mineral
1059
density and physical function in postmenopausal women with osteopenia and
1060
osteoporosis: the liftmor randomized controlled trial. Journal of Bone and Min-
1061
eral Research 33 (2), 211–220.
1062
Yokley, K., Tran, H. T., Pekari, K., Rappaport, S., Riihimaki, V., Rothman, N.,
1063
Waidyanatha, S., Schlosser, P. M., 2006. Physiologically-based pharmacoki-
63
1064
netic modeling of benzene in humans: a bayesian approach. Risk Analysis
1065
26 (4), 925–943.
1066
Yosibash, Z., Trabelsi, N., Milgrom, C., 2007. Reliable simulations of the human
1067
proximal femur by high-order finite element analysis validated by experimen-
1068
tal observations. Journal of Biomechanics 40 (16), 3688–3699.
1069
1070
Zaoui, A., 2002. Continuum micromechanics: survey. Journal of Engineering Mechanics 128 (8), 808–816.
64
Highlights:
Cellular micromechanics and pharmacokinetics integrated in a bone remodeling model Drug concentration at bone sites alters the dynamics of bone cell populations Osteoporosis and its treatment simulated in a mechanobiologic model Physiologically based pharmacokinetics model for bone seeking agents Bone tissue modeled by a permeability rate limited tissue compartment model
Ref: JMBBM_2019_1505 Title: A bone remodeling model governed by cellular micromechanics and physiologically based pharmacokinetics Journal: Journal of the Mechanical Behavior of Biomedical Materials
Declaration of interests X The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper. ☐The authors declare the following financial interests/personal relationships which may be considered as potential competing interests: