Skip to main content

Advertisement

Log in

A coupled mechano-biochemical model for bone adaptation

  • Published:
Journal of Mathematical Biology Aims and scope Submit manuscript

Abstract

Bone remodelling is a fundamental biological process that controls bone microrepair, adaptation to environmental loads and calcium regulation among other important processes. It is not surprising that bone remodelling has been subject of intensive both experimental and theoretical research. In particular, many mathematical models have been developed in the last decades focusing in particular aspects of this complicated phenomenon where mechanics, biochemistry and cell processes strongly interact. In this paper, we present a new model that combines most of these essential aspects in bone remodelling with especial focus on the effect of the mechanical environment into the biochemical control of bone adaptation mainly associated to the well known RANKL-RANK-OPG pathway. The predicted results show a good correspondence with experimental and clinical findings. For example, our results indicate that trabecular bone is more severely affected both in disuse and disease than cortical bone what has been observed in osteoporotic bones. In future, the methodology proposed would help to new therapeutic strategies following the evolution of bone tissue distribution in osteoporotic patients.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12
Fig. 13

Similar content being viewed by others

Notes

  1. In the biochemical model (Klika and Maršík 2010), the mechanical stimulation has different impacts on different considered reactions in order to reflect both inhibitory (e.g. osteoclastogenesis is inhibited by mechanical stimulation) and stimulatory (bone remodelling is stimulated overall by mechanical loading) effects of mechanical stimulation. That is why there are more mechanical parameters \(\mathcal {D}_{\alpha }\).

  2. The presented model can be extended with further details of biochemical control once considered necessary or advantageous.

Abbreviations

\(v_b\) :

Bone volume fraction [1]

\(v_m\) :

Mineral volume fraction of bone [1]

\(v_o\) :

Organic volume fraction of bone [1]

\(v_v\) :

Void volume fraction [1]

\(dam\) :

Damage concentration [1]

\(r_{\alpha }\) :

Rate of \(\alpha \)-th interaction \(\left[ \text{ kmol }\cdot \text{ m }^{-3}\cdot \text{ s }^{-1}\right] \)

\(\mathcal {A}_{\alpha }\) :

Affinity of \(\alpha \)-th reaction \(\left[ \text{ J }\cdot \text{ kmol }^{-1}\right] \)

:

The deformation rate tensor and its j-\(th\) invariant; \(j=1\) and \(j=2\) represent rate of volume dilatation and shear rate, respectively [\(s^{-1}\)]

\({\varvec{\varepsilon }},~\varepsilon ^{(j)}\) :

The deformation tensor \(\varepsilon \) and its j-\(th\) invariant [1]

\(l_{v\alpha },~l_{\alpha \alpha }\) :

Phenomenological constants in classical irreversible thermodynamics (CIT)

\(C_j, [\mathrm{C_j}]\) :

j-\(th\) substance and its molar concentration \(\left[ \text{ kmol }\cdot \text{ m }^{-3}\right] \)

\(\nu _{j\alpha },~\nu '_{j\alpha }\) :

Stoichiometric coefficients of substance \(C_j\) entering \(\alpha \)-th reaction or being produced in it (denoted with a prime), respectively [1]

\(k_{+\alpha },~k_{-\alpha }\) :

Forward and backward reaction rate constant of \(\alpha \)-th reaction

\(\alpha \) :

Ash fraction [1]

\(\nu \) :

Poisson ratio [1]

\(E\) :

Young’s modulus \(\left[ \text{ M } \text{ kg }\cdot \text{ m }^{-1}\cdot \text{ s }^{-1}\right] \)

\(\xi \) :

Daily strain history [1]

\(L\) :

Number of loading cases [1]

\(N_i\) :

Number of cycles of i-\(th\) loading case [1]

\(\bar{\varepsilon }\) :

Strain level [1]

\(\bar{d}\) :

Strain rate level [\(s^{-1}\)]

\(U\) :

Strain energy density \(\left[ \text{ kg }\cdot \text{ m }^{-1}\cdot \text{ s }^{-2}\right] \)

\(f_i\) :

Frequency of a considered loading case [\(s^{-1}\)]

\(T_i\) :

Period of a considered loading case [s]

\(\mathcal {D}_{\alpha }\) :

Parameters describing the influence of mechanical stimulus on \(\alpha -th\) interaction [1]

\(A\) :

Parameter used for relating \(\mathcal {D}_{\alpha }\) to \(\xi \) [1]

\(\mathcal {D}_{\alpha ,ref},~\xi _{ref}\) :

Reference values of mechanical stimulus [1]

\(\varepsilon _{crit}\) :

Critical strain value of bone tissue [1]

\(N_f\) :

Fatigue life expectation (number of cycles) [1]

\(K_i\) :

(\(i=t,c\)) Constant of proportionality in fatigue life expectation [1]

\(\delta _i\) :

(\(i=t,c\)) Exponent in fatigue life expectation [1]

\(dam_i\) :

(\(i=t,c\)) Damage in trabecular or cortical bone [1]

\(C_{c,t_1,t_2},~\gamma _{c,t}\) :

Fitted parameters in evolution laws for damage growth [1]

\(\varepsilon _u\) :

Ultimate tensile strain, function of calcium content [1]

\(v_{b,res}\) :

Amount of resorbed bone volume fraction [1]

\(\rho _{res}\) :

Difference in bone tissue density caused by resorption [kg\(\cdot \text{ m }^{-3}\)]

\(\rho _{max}\) :

Maximal amount of bone density (true bone density) [kg\(\cdot \text{ m }^{-3}\)]

\(J_{Old\_B}\) :

Normalised turnover rate in equilibrium [1]

\(n_{C_j}\) :

Normalised concentration of \(j\)-th substance [1]

\({n_{Old\_B}}_{,max}\) :

Maximal possible value of \(n_{Old\_B}\) corresponding to intact bone tissue under the maximal mechanical stimulus \(\xi (\varepsilon _{crit})\) [1]

\(m_m,~m_o\) :

Mineral/organic mass [kg\(\cdot \text{ m }^{-3}\)]

\(\rho _m,~\rho _o\) :

Mineral/organic true densities [kg\(\cdot \text{ m }^{-3}\)]

\(v_m^{\mathrm{secondary}}\) :

Exponential law for secondary mineralisation [1]

\(v_m^{prim}\) :

Mineral volume fraction value at the end of the primary phase [1]

\(v_j^0\) :

Initial value of \(v_j,~j=m,b,o\) [1]

\(\tau _{BR}\) :

Time corresponding to the end of the primary phase [1]

\(v_m^{max}\) :

\(v_m\) corresponding to the maximal mineral content found in bone [1]

\(\kappa \) :

Exponent in secondary mineralisation phase [1]

\(v_m^{\mathrm{OldB}}\) :

Average mineral content of old bone \(Old\_B\) [1]

\(v_m^{\mathrm{NewB}}\) :

Average mineral content of new bone \(New\_B\) [1]

\(n_{prim}\) :

Fraction of bone undergoing primary mineralisation [1]

\(X^0\) :

Initial value of \(X\) (e.g. \(v_b^0,\alpha ^0, [\mathrm{PTH}]_0\))

\(BT_{ind}\) :

Bone tissue index (a measure of total bone mass) [\(\text{ m }^3\)]

\(BMD_{ind}\) :

Bone mineral index (a measure of total mineral mass) [kg]

\(BMD(i)\) :

Bone mineral density at \(i\)-th element [kg\(\cdot \text{ m }^{-3}\)]

\(V(i)\) :

Volume of the \(i\)-th element [\(\text{ m }^3\)]

\(n_{C_{j,0}}\) :

Initial normalised concentration of j-\(th\) substance [1]

\([\mathrm{C_j}]_{st}\) :

Standard concentration of j-\(th\) substance [1]

\(n_{C_j,st}\) :

Normalised equivalent standard concentration of j-\(th\) substance [1]

\(n_{i_0}^{RRO,C_j}\) :

\(C_j \in \{PTH,~estr,~NO\}\), \(i \in \{RKL,~OPG\}\); predicted initial value of \(n_i\) (from a submodel for \(C_j\) influence on RANKL-RANK-OPG pathway) used as input in the RANKL-RANK-OPG model [1]

References

  • Aubin JE, Bonnelye E (2000) Osteoprotegerin and its ligand: a new paradigm for regulation of osteoclastogenesis and bone resorption. Medscape Womens Heal 5(2):5

    Google Scholar 

  • Beaupre GS, Orr TE, Carter DR (1990) An approach for time-dependent bone modeling and remodeling: theoretical development. J Orthop Res 8:651–661

    Article  Google Scholar 

  • Bergmann G, Graichen F, Rohlmann A (1993) Hip joint loading during walking and running, measured in two patients. J Biomech 26(8):969–990

    Article  Google Scholar 

  • Bougherara H, Klika V, Maršík F, Mařík I, Yahia L (2010) New predictive model for monitoring bone remodeling. J Biomed Mater Res Part A 95(1):9–24

    Article  Google Scholar 

  • Charopoulos I, Tournis S, Trovas G, Raptou P, Kaldrymides P, Skarandavos G, Katsalira K, Lyritis GP (2006) Effect of primary hyperparathyroidism on volumetric bone mineral density and bone geometry assessed by peripheral quantitative computed tomography in postmenopausal women. J Clin Endocr Metab 91(5):1748–1753

    Article  Google Scholar 

  • Cohen M Jr (2006) The new bone biology: pathologic, molecular, and clinical correlates. Am J Med Genet A 140(23):2646–2706

    Article  Google Scholar 

  • Collin-Osdoby P (2004) Regulation of vascular calcification by osteoclast regulatory factors RANKL and osteoprotegerin. Circ Res 95:1046–1057

    Article  Google Scholar 

  • Cowin SC, Hegedus DH (1976) Bone remodelling i: theory of adaptaive elasticity. J Elast 6:313–326

    Article  MATH  MathSciNet  Google Scholar 

  • Currey JD (2004) Tensile yield in compact bone is determined by strain, post-yield behaviour by mineral content. J Biomech 37(4):549–556

    Article  Google Scholar 

  • Doblaré M, García J (2002) Anisotropic bone remodelling model based on a continuum damage-repair theory. J Biomech 35(1):1–17

    Article  Google Scholar 

  • Doblaré M, García J, Gómez M (2004) Modelling bone tissue fracture and healing: a review. Eng Fract Mech 71:1809–1840

    Article  Google Scholar 

  • Ettinger B, Pressman A, Sklarin P, Bauer DC, Cauley JA, Cummings SR (1998) Associations between low levels of serum estradiol, bone density, and fractures among elderly women: The study of osteoporotic fractures. J Clin Endocr Metab 83(7):2239–2243

    Google Scholar 

  • Fang G, Ji B, Liu XS, Guo XE (2010) Quantification of trabecular bone microdamage using the virtual internal bond model and the individual trabeculae segmentation technique. Comput Method Biomech Biomed Eng 13(5):605–615

    Article  Google Scholar 

  • Frost HM (1963) Bone remodelling dynamics. C C Thomas, Springfield

    Google Scholar 

  • Frost HM (2000) The utah paradigm of skeletal physiology: an overview of its insights for bone, cartilage and collagenous tissue organs. J Bone Miner Metab 18:305–316

    Article  Google Scholar 

  • Fukunaga T, Kurata K, Matsuda J, Higaki H (2008) Effects of strain magnitude on mechanical responses of three-dimensional gel-embedded osteocytes studied with a novel 10-well elastic chamber. J Biomech Sci Eng 3(1):13–24

    Article  Google Scholar 

  • Fyhrie DP, Schaffler MB (1995) The adaptation of bone apparent density to applied load. J Biomech 28:135–146

    Article  Google Scholar 

  • García-Aznar JM, Rueberg T, Doblaré M (2005) A bone remodelling model coupling microdamage growth and repair by 3D BMU-activity. Biomech Model Mechanobiol 4:147–167

    Article  Google Scholar 

  • Goemaere S, Van Laere M, De Neve P, Kaufman J (1994) Bone mineral status in paraplegic patients who do or do not perform standing. Osteoporosis Int 4(3):138–143

    Article  Google Scholar 

  • Gong J, Arnold J, Cohn S (1964) Composition of trabecular and cortical bone. Anat Rec 149(3):325– 331

    Article  Google Scholar 

  • Hazelwood SJ, Martin RB, Rashid MM, Rodrigo JJ (2001) A mechanistic model for internal bone remodeling exhibits different dynamic responses in disuse and overload. J Biomech 34:299–308

    Article  Google Scholar 

  • Hazenberg JG, Taylor D, Lee TC (2007) The role of osteocytes and bone microstructure in preventing osteoporotic fractures. Osteoporosis Int 18:1–8

    Article  Google Scholar 

  • Hernandez C (2001) Simulation of bone remodeling during the development and treatment of osteoporosis. Phd thesis, Stanford University, Stanford

  • Hernández-Gil I, Gracia M, Pingarrón M, Jerez L (2006) Physiological bases of bone regeneration ii. the remodeling process. Med Oral Patol Oral Cir Bucal 11:E151–215

    Google Scholar 

  • Hill PA (1998) Bone remodelling. Brit J Orthod 25:101–107

    Article  Google Scholar 

  • Hjelmstad KD (2005) Fundamentals of structural mechanics. Springer, Berlin

    Google Scholar 

  • Huiskes R, Ruimerman R, van Lenthe GH, Janssen JD (2000) Effects of mechanical forces on maintenance and adaptation of form in trabecular bone. Nature 405:704–706

    Article  Google Scholar 

  • Ingber DE (2008) Tensegrity-based mechanosensing from macro to micro. Prog Biophys Mol Bio 97: 163179

    Article  Google Scholar 

  • Jacobs CR (1994) Numerical simulation of bone adaptation to mechanical loading. Phd thesis, Stanford University, Stanford

  • Jacobs CR, Simo JC, Beaupré GS, Carter DR (1997) Adaptive bone remodeling incorporating simultaneous density and anisotropy considerations. J Biomech 30(6):603–613

    Article  Google Scholar 

  • Jimi E, Akiyama S, Tsurukai T, Okahashi N, Kobayashi K, Udagawa N, Nishihara T, Takahashi N, Suda T (1999) Osteoclast differentiation factor acts as a multifunctional regulator in murine osteoclast differentiation and function. J Immunol 163(1):434–442

    Google Scholar 

  • Kim C, You L, Yellowley C, Jacobs C (2006) Oscillatory fluid flow-induced shear stress decreases osteoclastogenesis through rankl and opg signaling. Bone 39(5):1043–1047

    Article  Google Scholar 

  • Klika V (2010) Comparison of the effects of possible mechanical stimuli on the rate of biochemical reactions. J Phys Chem B 114(32):10,567–10,572

    Article  Google Scholar 

  • Klika V, Maršík F (2009) Coupling effect between mechanical loading and chemical reactions. J Phys Chem B 113:14,689–14,697

    Article  Google Scholar 

  • Klika V, Maršík F (2010) A thermodynamic model of bone remodelling: The influence of dynamic loading together with biochemical control. J Musculoskelet Neuron Interact 10(3):220–230

    Google Scholar 

  • Klika V, Maršík F (2011a) Biomechanics, vol 1, INTECH, Vienna, chap Feasible simulation of diseases related to bone remodelling and of their treatment. ISBN 978-953-307-312-5, [online] http://www.sciyo.com

  • Klika V, Maršík F (2011b) Feasible predictions of bone remodelling using modelling techniques. Locomot Appar 1+2:26–41, [online] http://www.pojivo.cz

  • Klika V, Maršík F, Mařík I (2010) Dynamic Modelling, INTECH, Vienna, chap Influencing the Effect of Treatment of Disease Related to Bone Remodelling by Dynamic Loading. ISBN 978-953-7619-68-8, [online] http://www.sciyo.com

  • Klika V, Grmela M (2013) Coupling between chemical kinetics and mechanics that is both nonlinear and compatible with thermodynamics. Phys rev E 87(1–1):012,141–012,141

    Article  Google Scholar 

  • Kobayashi Y, Udagawa N, Takahashi N (2009) Action of rankl and opg for osteoclastogenesis. Crit Rev Eukaryot Gene Expr 19(1):61

    Article  Google Scholar 

  • Komarova SV, Smith RJ, Dixon SJ, Sims SM, Wahl LM (2003) Mathematical model predicts a critical role for osteoclast autocrine regulation in the control of bone remodeling. Bone 33(2):206–215

    Article  Google Scholar 

  • Kroll MH (2000) Parathyroid hormone temporal effects on bone formation and resorption. Bull Math Biol 61(1):163–188

    Article  Google Scholar 

  • Kurata K, Heino TJ, Higaki H, Vaananen HK (2006) Bone marrow cell differentiation induced by mechanically damaged osteocytes in 3D gel-embedded culture. J Bone Miner Res 21(4):616–625

    Article  Google Scholar 

  • Lai Y, Qin L, Hung V, Chan K (2005) Regional differences in cortical bone mineral density in the weight-bearing long bone shafta pqct study. Bone 36(3):465–471

    Article  Google Scholar 

  • Lemaire V, Tobin FL, Greller LD, Cho CR, Suva LJ (2004) Modeling the interactions between osteoblast and osteoclast activities in bone remodeling. J Theor Biol 229(3):293–309

    Article  MathSciNet  Google Scholar 

  • Manolagas SC (1999) Cell number versus cell vigor-what really matters to a regeneration skeleton? Endocrinology 140(10):4377–4381

    Google Scholar 

  • Martin RB (1995) A mathematical model for fatigue damage repair and stress fracture in osteonal bone. J Orthop Res 13:309–316

    Google Scholar 

  • Martin RB (2007) Targeted bone remodeling involves BMU steering as well as activation. Bone 40(6):1574–1580

    Article  Google Scholar 

  • Martin T (2004) Paracrine regulation of osteoclast formation and activity: Milestones in discovery. J Musculoskelet Neuron Interact 4:243–253

    Google Scholar 

  • Martínez-Reina J, García-Aznar JM, Dominguez J, Doblaré M (2008) On the role of bone damage in calcium homeostasis. J Theor Biol 254:704–712

    Article  Google Scholar 

  • Nakashima T, Hayashi M, Fukunaga T, Kurata K, Oh-hora M, Feng J, Bonewald L, Kodama T, Wutz A, Wagner E, et al. (2011) Evidence for osteocyte regulation of bone homeostasis through rankl expression. Nat med 11:1231-1234

    Google Scholar 

  • van Oers RF, Ruimerman R, Tanck E, Hilbers PA, Huiskes R (2008) A unified theory for osteonal and hemiosteonal remodeling. Bone 42(2):250–259

    Article  Google Scholar 

  • Parfitt AM (2002) Life history of osteocytes: relationship to bone age, bone remodeling and bone fragility. J Musculoskelet Neuron Interact 2(6):499–500

    Google Scholar 

  • Pattin CA, Caler WE, Carter DR (1996) Cyclic mechanical property degradation during fatigue loading of cortical bone. J Biomech 29(1):69–79

    Article  Google Scholar 

  • Pivonka P, Zimak J, Smith D, Gardiner B, Dunstan C, Sims N, John Martin T, Mundy G (2008) Model structure and control of bone remodeling: a theoretical study. Bone 43(2):249–263

    Article  Google Scholar 

  • Pivonka P, Zimak J, Smith D, Gardiner B, Dunstan C, Sims N, John Martin T, Mundy G (2010) Theoretical investigation of the role of the rank-rankl-opg system in bone remodeling. J Theor Biol 262(2):306– 316

    Article  Google Scholar 

  • Prendergast PJ, Taylor D (1994) Prediction of bone adaptation using damage accumulation. J Biomech 27(8):1067–1076

    Article  Google Scholar 

  • Proff P, Römer P (2009) The molecular mechanism behind bone remodeling: a review. Clin Oral Invest 13:355–362

    Article  Google Scholar 

  • Qu C, Qin QH, Kang Y (2006) A hypothetical mechanism of bone remodeling and modeling under electromagnetic loads. Biomaterials 27:4050–4057

    Article  Google Scholar 

  • Ramtani S, Zidi M (2001) A theoretical model of the effect of continuum damage on a bone adaptation model. J Biomech 34(4):471–479

    Article  Google Scholar 

  • Rattanakul C, Lenbury Y, Krishnamara N, Wollkind DJ (2003) Modeling of bone formation and resorption mediated by parathyroid hormone: response to estrogen/PTH therapy. Biosystems 70:55–72

    Article  Google Scholar 

  • Robling AG, Castillo AB, Turner CH (2006) Biomechanical and molecular regulation of bone remodeling. Annu Rev Biomed Eng 8:455–498

    Article  Google Scholar 

  • Rodan G (1998) Bone homeostasis. Proc Natl Acad Sci USA 95(23):13,361–13,362

    Article  Google Scholar 

  • Rodan GA, Martin TJ (1981) Role of osteoblasts in hormonal control of bone resorption- a hypothesis. Calcif Tissue Int 33(4):349–351

    Article  Google Scholar 

  • Roux W (1881) Der zuchtende Kampf der teile, oder die Teilauslee im Organismus (Theorie der funktionellen anpassung). Wukgekn Ebgeknabb, Leipzig

    Google Scholar 

  • Rubin J, Murphy T, Nanes MS, Fan X (2000) Mechanical strain inhibits expression of osteoclast differentiation factor by murine stromal cells. Am J Physiol-Cell Ph 278:C1126–C1132

    Google Scholar 

  • Ryser M, Nigam N, Komarova S (2008) Mathematical modeling of spatio-temporal dynamics of a single bone multicellular unit. J Bone Miner Res 24(5):860–870

    Article  Google Scholar 

  • Sikavitsas VI, Temenoff JS, Mikos AG (2001) Biomaterials and bone mechanotransduction. Biomaterials 22:2581–2593

    Article  Google Scholar 

  • Skerry TM (1998) Methods in bone biology, 1st edn, chap 6, Chapman & Hall, London, UK, pp 149–176. ISBN 0 412 75770 2

  • Stacey E, Korkia P, Hukkanen M, Polak J, Rutherford O (1998) Decreased nitric oxide levels and bone turnover in amenorrheic athletes with spinal osteopenia. J Clin Endocr Metab 83(9):3056–3061

    Google Scholar 

  • Tan S, de Vries T, Kuijpers-Jagtman A, Semeins C, Everts V, Klein-Nulend J (2007) Osteocytes subjected to fluid flow inhibit osteoclast formation and bone resorption. Bone 41(5):745–751

    Article  Google Scholar 

  • Tudor-Locke C, Bassett J (2004) How many steps/day are enough?: Preliminary pedometer indices for public health. Sports Med 34(1):1–8

    Article  Google Scholar 

  • Vaira S, Alhawagri M, Anwisye I, Kitaura H, Faccio R, Novack DV (2008) RelA/p65 promotes osteoclast differentiation by blocking a RANKL-induced apoptotic JNK pathway in mice. J Clin Invest 118(6):2088–2097

    Google Scholar 

  • Virtama P, Telkkä A (1962) Cortical thickness as an estimate of mineral content of human humerus and femur. Brit J Radiol 35(417):632–633

    Article  Google Scholar 

  • Wang H, Ji B, Liu XS, Guo XE, Huang Y, Hwang KC (2012) Analysis of microstructural and mechanical alterations of trabecular bone in a simulated three-dimensional remodeling process. J biomech 45:2417-2425

    Google Scholar 

  • Wang H, Ji B, Liu X, Oers R, Guo X, Huang Y, Hwang KC (2013) Osteocyte-viability-based simulations of trabecular bone loss and recovery in disuse and reloading. Biomech Model Mechanobiol, pp 1–14. doi:10.1007/s10237-013-0492-1

  • Whalen RT, Carter DR, Steele CR (1988) Influence of physical activity on the regulation of bone density. J Biomech 21(10):825–837

    Article  Google Scholar 

  • Wimalawansa SJ (2007) Rationale for using nitric oxide donor therapy for prevention of bone loss and treatment of osteoporosis in humans. Ann NY Acad Sci 1117:283–297

    Article  Google Scholar 

  • Wolff J (1892) Das Gesetz der transformation der knochen. Hirchwild, Berlin

    Google Scholar 

Download references

Acknowledgments

This research was supported by the Instituto Aragones de Ciencias de la Salud through the research project (PIPAMER10/015) and the Spanish Ministry of Science and Technology through the Research Project DPI2011-22413. Further support was received from the CENTEM project, reg. no. CZ.1.05/2.1.00/03.0088, that is partially supported by the ERDF within the OP RDI Programme of the Ministry of Education, Youth and Sports and from the institutional support RVO:61388998.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Václav Klika.

Electronic supplementary material

Below is the link to the electronic supplementary material.

Supplementary material 1 (F 35 kb)

Supplementary material 2 (FOR 3 kb)

Appendices

Appendix A: Relation between daily strain history \(\xi \) and the mechanical parameter \(\mathcal {D}_{\alpha }\)

The mechanical stimulus used in the biochemical model was

$$\begin{aligned} \mathcal {D}_{\alpha }=l_{v\alpha }d^{(1)}, \end{aligned}$$
(23)

where \(d^{(1)}\) is the rate of volume variation and \(l_{v\alpha }\) are given in Table 8. In the mechano-biochemical model presented here, we generalise this relation by including influence of shear, see (6) and (7). The relation between the newly used form of mechanical stimulation \(\xi \) and the previous one \(\mathcal {D}_{\alpha }\) is the following:

$$\begin{aligned} \mathcal {D}_{\alpha }=A l_{v \alpha } \xi = \frac{\mathcal {D}_{\alpha ,ref}}{\xi _{ref}} \xi , \end{aligned}$$

where the value of \(A\) was determined from the reference healthy states, see the end of Sect. 2.2. The values of \(\mathcal {D}_{\alpha ,ref}\) are calculated from (23) where the value of \(d^{(1)}\) corresponds to the maximal strain rate found in a healthy mechanical stimulation, see Sect. 2.2 and the original model formulation (Klika and Maršík 2010):

$$\begin{aligned} \mathcal {D}_{1,ref}&= -4.58,\\ \mathcal {D}_{2,ref}&= 1.16\cdot 10^{-4},\\ \mathcal {D}_{3,ref}&= -1.31\cdot 10^{-2},\\ \mathcal {D}_{4,ref}&= -5.22\cdot 10^{-3},\\ \mathcal {D}_{5,ref}&= 2.04\cdot 10^{-4}. \end{aligned}$$

Appendix B: Biochemical model

The model presented here is an updated version of (Klika and Maršík 2010; Klika et al. 2010; Klika and Maršík 2011a, b). The core of the biochemical part of the presented model is characterised by the following interactions:

$$\begin{aligned} RR + MCELL&\mathop {\rightarrow }\limits ^{\mathrm{k}_{+1}}&MNOC + N_4,\nonumber \\ MNOC + Old\_B&\mathop {\rightarrow }\limits ^{\mathrm{k}_{+2}}&LF + N_7,\nonumber \\ LF + osteoprogenitor&\mathop {\rightarrow }\limits ^{\mathrm{k}_{+3}}&OB + N_{10},\nonumber \\ OB + N_{11}&\mathop {\rightarrow }\limits ^{\mathrm{k}_{+4}}&Osteoid + N_{13},\nonumber \\ N_{14} + Osteoid&\mathop {\rightarrow }\limits ^{\mathrm{k}_{+5}}&New\_B + N_{16}, \end{aligned}$$
(24)

where \(MNOC\) are multi-nucleated osteoclasts being formed by fusion from their precursors \(MCELL\) when appropriate receptor-ligand (cell-to-cell) interaction takes place (\(RR\) representing RANKL-RANK interaction mediated by the RANKL-RANK-OPG chain), \(Old\_B\) denotes old bone that might be resorbed, the released local factors from bone are denoted by \(LF\) which activate \(osteoprogenitor\) cells that proliferate and differentiate into osteoblasts \(OB\), which secrete \(osteoid\) (non-mineralised bone - organic part of bone tissue). The ossification of \(osteoid\) into new bone tissue \(New\_B\) happens after addition of appropriate substances, \(N_{14}\), takes place. \(N_4, N_7, N_{10}, N_{13}, N_{16}\) are waste products. For a more detailed description see (Klika and Maršík 2010).

The RANKL-RANK-OPG pathway is considered as follows:

$$\begin{aligned}&RANKL + RANK~\mathop {\rightleftarrows }\limits ^{\mathrm{k}_{\pm 1}} RR, \end{aligned}$$
(25)
$$\begin{aligned}&\quad RANKL + OPG~\mathop {\rightleftarrows }\limits ^{\mathrm{k}_{\pm 2}} RO_{\mathrm{inactive}}, \end{aligned}$$
(26)

Finally, the effect of estradiol, PTH, and NO is translated through this pathway:

$$\begin{aligned} Estradiol + OPG_{\mathrm{prod}}+Subst&\mathop {\rightleftarrows }\limits ^{\mathrm{k}_{\pm 1}}&OPG+OPG_{\mathrm{prod}},\end{aligned}$$
(27)
$$\begin{aligned} PTH + RANKL_{\mathrm{prod}}+Subst&\mathop {\rightleftarrows }\limits ^{\mathrm{k}_{\pm 1}}&RANKL+RANKL_{\mathrm{prod}},\end{aligned}$$
(28)
$$\begin{aligned} NO + RANKL~&\mathop {\rightleftarrows }\limits ^{\mathrm{k}_{\pm 1}}&Remaining\_prod,\end{aligned}$$
(29)
$$\begin{aligned} NO + OPG_{\mathrm{prod}}+Subst&\mathop {\rightleftarrows }\limits ^{\mathrm{k}_{\pm 2}}&OPG+OPG_{\mathrm{prod}}, \end{aligned}$$
(30)

where \(OPG_{\mathrm{prod}},~RANKL_{\mathrm{prod}}\) represent the group of cells that expresses OPG, RANKL and a mixture of substances needed for osteoprotegerin, RANKL production is noted as \(Subst\), respectively and \(Remaining\_prod\) represents inactivated RANKL.

The dimensionless system of differential equations describing the kinetics of interactions of the core of the biochemical part of the presented model model (24) together with the influence of mechanical stimulus \(\mathcal {D}_{\alpha }(\xi )\) (for the relationship between \(\mathcal {D}_{\alpha }\) and \(\xi \) see appendix 5) is provided here

$$\begin{aligned} \frac{\mathrm{d}n_{MCELL}}{\mathrm{d}\tau }&= -\delta _1(\beta _1+n_{MCELL})n_{MCELL}+\mathcal {J}_3\nonumber \\&+\mathcal {J}_{New\_B}-\mathcal {D}_1(\xi )\\ \frac{\mathrm{d}n_{Old\_B}}{\mathrm{d}\tau }&= -(\beta _3-n_{MCELL}+n_{Old\_B})n_{Old\_B}-\mathcal {D}_2(\xi )\nonumber \\&+\mathcal {J}_{New\_B} \\ \frac{\mathrm{d}n_{OB}}{\mathrm{d}\tau }&= \delta _3(\beta _6-n_{Old\_B}-(n_{OB}+n_{Osteoid}+n_{New\_B}))\nonumber \\&\times (\beta _8-(n_{OB}+n_{Osteoid}+n_{New\_B}))\\&-\delta _4(\beta _{11}-(n_{Osteoid}+n_{New\_B}))n_{OB}+\mathcal {D}_3(\xi )-\mathcal {D}_4(\xi ) \nonumber \\ \frac{\mathrm{d}n_{Osteoid}}{\mathrm{d}\tau }&= \delta _4(\beta _{11}-(n_{Osteoid}+n_{New\_B}))n_{OB}\nonumber \\&-\delta _5 (\beta _{14}-n_{New\_B})n_{Osteoid}+\mathcal {D}_4(\xi )-\mathcal {D}_5(\xi ) \\ \frac{\mathrm{d}n_{New\_B}}{\mathrm{d}\tau }&= \delta _5 (\beta _{14} -n_{New\_B})n_{Osteoid}-\mathcal {J}_{New\_B}+\mathcal {D}_5(\xi ), \end{aligned}$$

where \(\beta _{i}\) is the sum of normalised initial molar concentrations of relevant substances, \(\delta _{\alpha }\) relates to the interaction rate, \(D_{\alpha }(\xi )\) describes the influence of dynamic loading on reactions, \(n_{i}\) is the normalised concentration of the \(i\)-th substance, \(\mathcal {J}_3\) is a normalised outflow of \(MNOC\) (i.e., \(MNOC\) apoptosis) and \(\mathcal {J}_{New\_B}\) is the normalised outflow of new bone \(New\_B\) into old bone \(Old\_B\).

By solving these kinetic equations, the time evolution of \([\mathrm{MCELL}], [\mathrm{Old\_B}], [\mathrm{OB}], [\mathrm{Osteoid}]\), and \([\mathrm{New\_B}]\) concentrations can be obtained: the relation between concentrations and their dimensionless form is:

$$\begin{aligned} n_{N_i}=[\mathrm{N_i}] \frac{1}{k_{+2} [\mathrm{Bo}]^2}, \end{aligned}$$

where \(k_{+2}=6 \cdot 10^7 \frac{l}{mol\cdot s}\) and \([\mathrm{Bo}]=2.2 \frac{10^{-4}}{N_A} \frac{mol}{\mu m^3}\) with \(N_A\) being the Avogardo’s number. In this model, it is assumed that the density of bone tissue is proportional to the number of osteocytes in a given volume, or that the molar concentration of osteocytes in tissue is constant; the amount of bone tissue in a given volume then determines the local bone density. Thus the quantity \([\mathrm{Old\_B}]\) is (by definition) equal to the concentration of osteocytes \([\mathrm{OCy}]\) which, with respect to above mentioned assumption, gives us the concentration of bone tissue \(Old\_B\):

$$\begin{aligned}{}[\mathrm{Old\_B}]=\frac{2.3}{0.81} n_{Old\_B} \left[ \frac{kg}{m^3}\right] \end{aligned}$$

which directly relates the maximal normalised \(n_{Old\_B}\) to maximal apparent bone tissue density \([\mathrm{Old\_B}]\) found in the human femur. The kinetics of other interactions is derived from the Law of Mass Action or can be found in the referred literature. Here, to provide a stand-alone model, the remaining differential equations of the biochemical model describing the considered kinetics of RANKL-RANK-OPG pathway are recapitulated.

RANKL-RANK-OPG pathway The law of mass action gives for the RANKL-RANK-OPG pathway (25) the following system of differential equations where we denote the dimensionless concentration of RANKL and OPG as \(n_{RKL}\) and \(n_{OPG}\), respectively:

$$\begin{aligned} \frac{\mathrm{d}n_{RKL}}{\mathrm{d}\tau }&= -n_{RKL}\left( \beta _{\mathrm{RK}}^{\mathrm{RRO}}+n_{RKL}-n_{OPG}\right) \nonumber \\&+\delta _{-1}^{\mathrm{RRO}}\left( \beta _{\mathrm{RR}}^{\mathrm{RRO}}-n_{RKL}+n_{OPG}\right) \nonumber \\&-{}\delta _{+2}^{\mathrm{RRO}}n_{RKL}n_{OPG}+\delta _{-2}^{\mathrm{RRO}}\left( \beta _{\mathrm{RO}}^{\mathrm{RRO}}-n_{OPG}\right) ,\\ \frac{\mathrm{d}n_{OPG}}{\mathrm{d}\tau }&= -\delta _{+2}^{\mathrm{RRO}}n_{RKL}n_{OPG}+\delta _{-2}^{\mathrm{RRO}}\left( \beta _{\mathrm{RO}}^{\mathrm{RRO}}-n_{OPG}\right) ,\nonumber \end{aligned}$$

with \(n_{RR}(t)=\beta _{\mathrm{RR}}^{\mathrm{RRO}}-n_{RKL}(t)+n_{OPG}(t)\) and parameter values are given in Appendix 8. The amount of RANKL-RANK bonds (\(RR\)) mediate the outcome effect of the whole RANKL-RANK-OPG pathway on bone remodelling. The amount of \(RR\) bonds is included in the \(\beta _1\) parameter. The differential equations are in dimensionless form and to provide prediction of changes in real serum levels they need to be compared with the physiological value of RANKL that was used for obtaining the dimensionless form. In total, the initial conditions are given by relationships

$$\begin{aligned} n_{RKL_0}&=\frac{[\mathrm{RKL}]}{[\mathrm{RKL}]_{st}}\underbrace{n_{RKL,st}}_{1},\\ n_{OPG_0}&=\frac{[\mathrm{OPG}]}{[\mathrm{OPG}]_{st}}\underbrace{n_{OPG,st}}_{0.2333}, \end{aligned}$$

and standard serum levels in humans were estimated to be \([\mathrm{RKL}]_{st}=0.84\frac{pmol}{l}=46.2\frac{pg}{ml}\) and \([\mathrm{OPG}]_{st}=1.8\frac{pmol}{l}=36\frac{pg}{ml}\). The normalised standard OPG concentration \(n_{OPG,st}\) is determined from molar masses of RANKL and OPG, see (Klika et al. 2010). Finally, the interconnection with the core model is

$$\begin{aligned} \beta _1=(1.41/0.79) n_{RR}\left( \tau _{7days}^{RRO}\right) - 0.81, \end{aligned}$$

giving \(\beta _1=0.6\) for standard serum values of RANKL and OPG. For more details see (Klika et al. 2010).

As was mentioned, many of the factors having a significant influence on RANKL-RANK-OPG signalling pathway, and thus on bone quality, are translated through this same pathway. Below we will add impacts of PTH, estradiol and nitric oxide. Here, we model their impact by modifying the initial values of RANKL and OPG in this RANKL-RANK-OPG model as predicted from the following sub-models. The final initial conditions for RANKL and OPG are determined by the sum of all predicted deflections from their standard values, namely

$$\begin{aligned} n_{RKL_0}&= n_{RKL,st}+\left( n_{RKL_0}^{RRO,PTH}-n_{RKL,st}\right) + \left( n_{RKL_0}^{RRO,NO}-n_{RKL,st}\right) ,\qquad \end{aligned}$$
(31)
$$\begin{aligned} n_{OPG_0}&= n_{OPG,st}+\left( n_{OPG_0}^{RRO,estr}-n_{OPG,st}\right) + \left( n_{OPG_0}^{RRO,NO}-n_{OPG,st}\right) ,\qquad \end{aligned}$$
(32)

where the terms in brackets denote differences from standard initial values caused by the control mechanism noted in the upper indices.

Further, as each following sub-model has its own normalisation there are two scaling factors present in each one: (1) scaling of used in-vitro data by \(C_{in-vitro}\) (it is unclear if this scaling finds an equivalent amount, in terms of its effects on bone remodelling, of in-vivo serum levels to that of used in-vitro ones but it seems to capture their impact on bone remodelling as reported in literature), (2) scaling to match the different normalisation used— matching is such that standard value of a given factor are met in both normalisations by \(C_{norm}\), see below.

PTH The impact of PTH on RANKL-RANK-OPG pathway (28) can be described as follows:

$$\begin{aligned} \frac{\mathrm{d}{\tilde{n}_{PTH}}}{\mathrm{d}\tau }=-{\tilde{n}_{PTH}}\left( \beta _{\mathrm{Substr}}^{\mathrm{PTH}}+{\tilde{n}_{PTH}}\right) +\delta _{-1}^{\mathrm{PTH}}\left( \beta _{\mathrm{RKL}}^{\mathrm{PTH}}-{\tilde{n}_{PTH}}\right) , \end{aligned}$$

which affects the amount of RANKL available for making receptor-ligand bonds. This equation for normalised in-vitro concentration \(\tilde{n}_{PTH}\) was fitted with in-vitro dose-dependent data (no other available). The rescaling of in-vitro data is included in \({\tilde{n}_{PTH}}=C_{in-vitro}^{PTH} n_{PTH}\), \(C_{in-vitro}^{PTH}=0.296\) and thus the initial condition needed for solving the above differential equation is

$$\begin{aligned} {\tilde{n}_{PTH_0}}=C_{in-vitro}^{PTH} \frac{[\mathrm{PTH}]_0}{[\mathrm{PTH}]_{st}}, \end{aligned}$$
(33)

where \([\mathrm{PTH}]\) represents serum levels of PTH, typically in \(\frac{pg}{ml}\). The concentration of RANKL after action of PTH is (\(n_{PTH}(\tau )\) being the scaled solution of the above differential equation):

$$\begin{aligned} n_{RKL_0}^{RRO,PTH}&= C_{norm}^{PTH}\left( \beta _{\mathrm{RKL}}^{\mathrm{PTH}}-n_{PTH}\left( \tau _{10days}^{PTH}\right) \right) \\ {}&= C_{norm}^{PTH}\left( n_{RKL_0}^{PTH}+n_{PTH_0}-n_{PTH}\left( \tau _{10days}^{PTH}\right) \right) , \end{aligned}$$

where \(n_{RKL_0}^{PTH}\) represents the initial concentration of RANKL before interacting with PTH according to (28) and \(C_{norm}^{PTH}=1.0\) is the second scaling to match the different normalisations (see above). The amount of available RANKL for forming bonds with its receptor RANK is denoted as \(n_{RKL_0}^{RRO,PTH}\) and is used as an input into the RANKL-RANK-OPG model taking into account the influence of PTH on RANKL concentration, see (31)–(32). Parameter values can be found again in Appendix 8.

Estradiol The idea behind adding an impact of estradiol on RANKL-RANK-OPG pathway (27) is the same as in PTH, only that estradiol affects OPG instead of RANKL (over-tilde has the same meaning as in PTH—rescaled values of estradiol):

$$\begin{aligned} \frac{\mathrm{d}{\tilde{n}_{estr}}}{\mathrm{d}\tau }=-{\tilde{n}_{estr}}(\beta _{\mathrm{Substr}}^{\mathrm{estr}}+{\tilde{n}_{estr}})+ \delta _{-1}^{\mathrm{estr}}\left( \beta _{\mathrm{OPG}}^{\mathrm{estr}}-{\tilde{n}_{estr}}\right) , \end{aligned}$$

and the initial condition is obtained from serum levels analogously to that one in PTH, Eq. (33), with \(C_{in-vitro}^{estr}=1.19\). The concentration of OPG after action of estradiol is

$$\begin{aligned} n_{OPG_0}^{RRO,estr}&= C_{norm}^{estr}\left( \beta _{\mathrm{OPG}}^{\mathrm{estr}}-n_{estr}\left( \tau _{\mathrm{24h}}^{\mathrm{estr}}\right) \right) \\ {}&= C_{norm}^{estr}\left( n_{OPG_0}^{estr}+n_{estr_0}-n_{estr}\left( \tau _{\mathrm{24h}}^{\mathrm{estr}}\right) \right) , \end{aligned}$$

with \(C_{norm}^{estr}=1.04\). Again the amount of available OPG for blocking bond formation is denoted as \(n_{OPG_0}^{RRO,estr}\) and is used as an input into the RANKL-RANK-OPG model taking into account the influence of estradiol on OPG concentration, see (31)-(32). Parameter values can be found again in Appendix 8.

Nitric oxide Similarly, the influence of nitric oxide on RANKL-RANK-OPG pathway (29)–(30) can be included. It affects both RANKL and OPG concentrations. It was assumed that the effects of NO on OPG and RANKL can be separated (Klika and Maršík 2011a) leading to the same differential equations as in previous cases for estradiol and PTH. Due to unavailability of in-vivo data that could quantify NO effects on bone quality we used studies where the amount of NO was regulated by NO donor intake in milligrams per day (typically nitroglycerin). A non-linear scaling in the input data was required to obtain a fit with observed outcome effects on BMD. It is unclear if this scaling finds an equivalent amount (in terms of its effects on bone remodelling) of in-vivo NO serum levels to that of donor but it seems to capture the impact of NO on bone remodelling. Concretely, the scaling transforms were logarithmic \({\tilde{n}_{NO,OPG}}=C_{in-vitro}^{NO,OPG}\left( 1+\ln \left( [\mathrm{NO^{OPG}}]/[\mathrm{NOdonor}]_{stand}\right) \right) \) and \({\tilde{n}_{NO^{RKL}}}=C_{in-vitro}^{NO,RKL}\left( 1+\ln \left( [\mathrm{NO^{RKL}}]/[\mathrm{NOdonor}]_{stand}\right) \right) \) with normalisation constants \(C_{in-vitro}^{NO,OPG}=1.01\), \(C_{in-vitro}^{NO,RKL}=0.036\) and with \([\mathrm{NOdonor}]_{stand}=0.044\frac{mg}{kg\times day}\) being the standard or reference amount of NO donor that does not cause any change in BMD (Wimalawansa 2007).

Finally, the differences from standard values of OPG and RANKL caused by NO are

$$\begin{aligned} n_{RKL_0}^{RRO,NO}&= C_{norm}^{NO,RKL}\left( \beta _{\mathrm{RKL}}^{\mathrm{NO,RKL}}-n_{NO,RKL}\left( \tau _{\mathrm{24h}}^{\mathrm{NO,RKL}}\right) \right) \\&= C_{norm}^{NO,RKL}\left( n_{OPG_0}^{NO,RKL}+n_{NO,RKL_0}-n_{NO,RKL}\left( \tau _{\mathrm{24h}}^{\mathrm{NO,RKL}}\right) \right) ,\\ n_{OPG_0}^{RRO,NO}&= C_{norm}^{NO,OPG}\left( \beta _{\mathrm{OPG}}^{\mathrm{NO,OPG}}-n_{NO,OPG}\left( \tau _{\mathrm{24h}}^{\mathrm{NO,OPG}}\right) \right) \\&= C_{norm}^{NO,OPG}\left( n_{OPG_0}^{NO,OPG}+n_{NO,OPG_0}-n_{NO,OPG}\left( \tau _{\mathrm{24h}}^{\mathrm{NO,OPG}}\right) \right) , \end{aligned}$$

where the two normalisation constants are \(C_{norm}^{NO,RKL}=1.02\), \(C_{norm}^{NO,OPG}=2.16\). Again the amount of available OPG for blocking bond formation is denoted as \(n_{OPG_0}^{RRO,NO},~n_{RKL_0}^{RRO,NO}\) and is used as an input into the RANKL-RANK-OPG model taking into account the influence of nitric oxide on this essential control pathway, see (31)–(32). Parameter values can be found in Appendix 8.

Implementation of the whole model in Abaqus as a user subroutine is enclosed in the supplementary material.

Appendix C: Stationary solution

The biochemical model, see Sect. 6, has exactly one positive stationary solution:

$$\begin{aligned} \overline{n_{MCELL}}&= 1/2\Bigg (-\beta _1+\sqrt{\beta _1^2+4\frac{-\mathcal {D}_1(\xi )+\mathcal {J}_3+\mathcal {J}_{14}}{\delta _1}}\Bigg )\nonumber \\ \overline{n_{Old\_B}}&= 1/2\Bigg (\bigg (-\beta _3+\overline{n_{MCELL}}\bigg )\nonumber \\&+\sqrt{\bigg (-\beta _3+\overline{n_{MCELL}}\bigg )^2+4(\mathcal {J}_{14}-\mathcal {D}_2(\xi ))}\Bigg )\nonumber \\ Sum_1&= 1/2\Bigg (\bigg (\beta _8+\beta _6-\overline{n_{Old\_B}}\bigg )\\&{}-\sqrt{\bigg (\beta _8-(\beta _6-\overline{n_{Old\_B}})\bigg )^2+4\frac{\mathcal {J}_{14}-\mathcal {D}_3(\xi )}{\delta _3}}\Bigg ) \nonumber \\ Sum_2&= 1/2\Bigg (\bigg (\beta _{11}+Sum_1\bigg )-\sqrt{\bigg (\beta _{11}-Sum_1\bigg )^2+4\frac{\mathcal {J}_{14}-\mathcal {D}_4(\xi )}{\delta _4}}\Bigg ) \nonumber \\ \overline{New\_B}&= 1/2\Bigg (\bigg (\beta _{14}+Sum_2\bigg )-\sqrt{\bigg (\beta _{14}-Sum_2\bigg )^2+4\frac{\mathcal {J}_{14}-\mathcal {D}_5(\xi )}{\delta _5}}\Bigg ) \nonumber \\ \overline{n_{Osteoid}}&= Sum_2-\overline{New\_B}\nonumber \\ \overline{n_{OB}}&= Sum_1-Sum_2.\nonumber \end{aligned}$$

which for the considered parameter values fulfil all the assumptions of Poincare-Ljapunov theorem and thus the stationary solution is an asymptotically stable fixed point.

Appendix D: List of parameter values

$$\begin{aligned}&\text{ initial } \text{ values } \text{ of } \text{ damage, } v_b, \alpha \text{(ash } \text{ fraction) }\\&\quad d_0=0.0, v_b^0=0.6, \alpha _0=0.69\\&\mathbf{Biochemical}\,\, \mathbf{part}\,\, \mathbf{of}\,\, \mathbf{the}\,\, \mathbf{model}\\&\quad [\mathrm{PTH}]_{st}=34.0\frac{pg}{ml}, [\mathrm{estr}]_{st}=50.0\frac{pg}{ml},[\mathrm{NO_{donor}}]_{st}=0.044\frac{mg}{kg\cdot day},\\&\qquad [\mathrm{RKL}]_{st}=46.2\frac{pg}{ml},[\mathrm{OPG}]_{st}=36.0\frac{pg}{ml}\\&\text{ Estradiol } \text{ submodel }\\&\quad \tau _{24h}^{estr}=26.166,\delta _1^{estr}=0.1448, nSUBST_0^{estr}=0.018\\&\text{ NO-OPG } \text{ submodel }\\&\quad \tau _{24h}^{NO,OPG}=0.00177,\delta _1^{NO,OPG}=476.7, nSUBST_0^{NO,OPG}=7.67\\&\text{ NO-RANKL } \text{ submodel }\\&\quad \tau _{24h}^{NO,RKL}=50.0, \delta _1^{NO,RKL}=0.8, nREMAIN_0^{NO,RKL}=1E-5\\&\text{ PTH } \text{ submodel }\\&\quad \tau _{10days}^{PTH}=1687.3, \delta _1^{PTH}=1.763, nSUBST_0^{PTH}=2.6,\\&\quad nRKL_0^{PTH}=1.31\\&\text{ RANKL-RANK-OPG } \text{ submodel }\\&\quad \tau _{7days}^{RRO}=4.64,\delta _1^{RRO}=4.92E-6, \delta _{2+}^{RRO}=12.96,\\&\quad \delta _{2-}^{RRO}=5.86E-19,RKL_{Stand}=1.0,OPG_{Stand}=0.2333\\&\text{ main } \text{ model }\\&\quad \tau _{Day}=1 \text{ day },\tau _{BR}=315 \text{ days }\\&\quad MCELL_0=0.8,osteoprog_0=0.0,Osteoid_0=0.0,\\&\quad \delta _1=4.0,\delta _3=2.0,\delta _4=0.142857,\delta _5=0.142857,\\&\quad \beta _6=1.05,\beta _8=0.1,J_{Old\_B}=2.6E-4,J_3=4/25 J_{Old\_B},\\&\quad \beta _{11}=1.0,\beta _3=0.0005,\\&\quad \beta _{14}=0.05,\mathcal {D}_1^{ref}=-4.58,\mathcal {D}_2^{ref}=1.16E-4,\mathcal {D}_3^{ref}=-1.31E-2,\\&\quad \mathcal {D}_4^{ref}=-5.22E-3,\mathcal {D}_5^{ref}=2.04E-4 \end{aligned}$$

Rights and permissions

Reprints and permissions

About this article

Cite this article

Klika, V., Pérez, M.A., García-Aznar, J.M. et al. A coupled mechano-biochemical model for bone adaptation. J. Math. Biol. 69, 1383–1429 (2014). https://doi.org/10.1007/s00285-013-0736-9

Download citation

  • Received:

  • Revised:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00285-013-0736-9

Keywords

Mathematics Subject Classification (2000)

Navigation