Skip to main content
Log in

Mechanistic modeling of vascular tumor growth: an extension of Biot’s theory to hierarchical bi-compartment porous medium systems

  • Original Paper
  • Published:
Acta Mechanica Aims and scope Submit manuscript

Abstract

Existing continuum multiphase tumor growth models typically do not include microvasculature, or if present, this is modeled as a non-deformable network of vessels. Vasculature behavior and blood flow are usually non-coupled with the underlying tumor phenomenology from the mechanical viewpoint; hence, phenomena like vessel compression/occlusion modifying microcirculation and oxygen supply cannot be taken into account. Here, the tumor tissue is modeled as a reactive bi-compartment porous medium: the extracellular matrix constitutes the solid scaffold; blood flows in the vascular porosity, whereas the extravascular porous compartment is saturated by two cell phases and interstitial fluid (mixture of water and nutrient species). The pressure difference between blood and the extravascular overall pressure is sustained by vessel walls and drives shrinkage or dilatation of the vascular porosity. Model closure is achieved thanks to a consistent non-conventional definition of the Biot’s effective stress tensor. Angiogenesis is modeled by introducing a vascularization state variable and accounting for tumor angiogenic factors and endothelial cells. Closure relationships and mass exchange terms related to vessel formation are detailed in a numerical example reproducing the principal features of angiogenesis. This example is preceded by a first pedagogical numerical study on one-dimensional bio-consolidation. Results demonstrate that the bi-compartment poromechanical model is fully coupled (the external loads impact fluid flow in both porous compartments) and that it can serve as a basis for further applications like modeling of drug delivery and tissue ulceration.

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
Fig. 14

Similar content being viewed by others

References

  1. Michor, F., Liphardt, J., Ferrari, M., Widom, J.: What does physics have to do with cancer? Nat. Rev. Cancer 11(9), 657–670 (2013)

    Article  Google Scholar 

  2. Dusheck, J.: Oncology: getting physical. Nature 491, S50–S51 (2012)

    Article  Google Scholar 

  3. Jonietz, E.: Mechanics: the forces of cancer. Nature 491(7425), S56 (2012)

    Article  Google Scholar 

  4. Ferrari, M.: Frontiers in cancer nanomedicine: directing mass transport through biological barriers. Trends Biotechnol. 28(4), 181–8 (2010)

    Article  Google Scholar 

  5. Jain, R.K., Martin, J.D., Stylianopoulos, T.: The role of mechanical forces in tumor growth and therapy. Annu. Rev. Biomed. Eng. 16, 321–46 (2014)

    Article  Google Scholar 

  6. Vilanova, G., Colominas, I., Gomez, H.: A mathematical model of tumour angiogenesis: growth, regression and regrowth. J. R. Soc. Interface 14, 20160918 (2017)

    Article  Google Scholar 

  7. Santagiuliana, R., Ferrari, M., Schrefler, B.: Simulation of angiogenesis in a multiphase tumor growth model. Comput. Methods Appl. Mech. Eng. 304, 197 (2016)

    Article  MathSciNet  Google Scholar 

  8. Kremheller, J., Vuong, A.T., Schrefler, B., Wall, W.: An approach for vascular tumor growth based on a hybrid embedded/homogenized treatment of the vasculature within a multiphase porous medium model. Int. J. Numer. Methods Biomed. Eng. 35, e3253 (2019)

    Article  MathSciNet  Google Scholar 

  9. Scianna, M., Bell, C.G., Preziosi, L.: A review of mathematical models for the formation of vascular networks. J. Theor. Biol. 333, 174 (2013)

    Article  MathSciNet  Google Scholar 

  10. Sciumè, G., Gray, W.G., Ferrari, M., Decuzzi, P., Schrefler, B.A.: On computational modeling in tumor growth. Arch. Comput. Methods Eng. 20(4), 327–52 (2013)

    Article  MathSciNet  Google Scholar 

  11. Gray, W.G., Miller, C.T.: Introduction to the Thermodynamically Constrained Averaging Theory for Porous Medium Systems. Springer, Berlin (2014)

    Book  Google Scholar 

  12. Sciumè, G., Shelton, S., Gray, W.G., Miller, C.T., Hussain, F., Ferrari, M., Decuzzi, P., Schrefler, B.: A multiphase model for three-dimensional tumor growth. New J. Phys. 15, 015005 (2013)

    Article  Google Scholar 

  13. Sciumè, G., Ferrari, M., Schrefler, B.A.: Saturation-pressure relationships for two- and three-phase flow analogies for soft matter. Mech. Res. Commun. 62, 132 (2014)

    Article  Google Scholar 

  14. Sciumè, G., Gray, W.G., Hussain, F., Ferrari, M., Decuzzi, P., Schrefler, B.A.: Three phase flow dynamics in tumor growth. Comput. Mech. 53(3), 465 (2014)

    Article  MathSciNet  Google Scholar 

  15. Sciumè, G., Santagiuliana, R., Ferrari, M., Decuzzi, P., Schrefler, B.A.: A tumor growth model with deformable ECM. Phys. Biol. 11(6), 065004 (2014)

    Article  Google Scholar 

  16. Bearer, E.L., Lowengrub, J., Frieboes, H.B., Chuang, Y.L., Jin, F., Wise, S.M., Ferrari, M., Agus, D.B., Cristini, V.: Multiparameter computational modeling of tumor invasion. Cancer Res. 69(10), 4493–501 (2009)

    Article  Google Scholar 

  17. Sciumè, G.: A hierarchical multi-compartment porous medium system for evolution of tumor microenvironment during avascular and vascular growth. In: ALERT Workshop 2017 (ALERT GEOMATERIALS—The Alliance of Laboratories in Europe for Education, Research and Technology, 2017). http://alertgeomaterials.eu/presentations-of-the-alert-workshop-2017/ (2017)

  18. Sciumè, G.: A two-compartment hierarchical porous medium system for vascular tumor growth: theory and implementation in Cast3M. In: Club Cast3M 2017 (CEA— French Atomic Energy and Alternative Energy Commission, 2017). http://www-cast3m.cea.fr/index.php?xml=clubcast3m2017 (2017)

  19. Kremheller, J., Vuong, A.T., Yoshihara, W., Schrefler, B., Wall, W.: A monolithic multiphase porous medium framework for (a-)vascular tumor growth. Comput. Methods Appl. Mech. Eng. 340, 657 (2018)

    Article  MathSciNet  Google Scholar 

  20. Lewis, R.W., Schrefler, B.A.: The Finite Element Method in the Static and Dynamic Deformation and Consolidation of Porous Media, 2nd, Edition edn. Wiley, Hoboken (1998)

    MATH  Google Scholar 

  21. Le Maout, V.D., Alessandri, K., Gurchenkov, B., Bertin, h, Nassoy, P., Sciumè, G.: Role of mechanical cues and hypoxia on the growth of tumor cells in strong and weak confinement: a dual in vitro–in silico approach. Sci. Adv. 6, eaaz7130 (2020)

    Article  Google Scholar 

  22. Anderson, A.R.A., Chaplain, M.A.J.: Continuous and discrete mathematical models of tumor-induced angiogenesis. Bull. Math. Biol. 60, 857–899 (1998)

    Article  Google Scholar 

  23. Anderson, A., Chaplain, M.: Continuous and discrete mathematical models of tumor-induced angiogenesis. Bull. Math. Biol. 60(5), 857 (1998)

    Article  Google Scholar 

  24. Heldin, C.H., Rubin, K., Pietras, K., Ostman, A.: High interstitial fluid pressure: an obstacle in cancer therapy. Nat. Rev. Cancer 10, 806–813 (2004)

    Article  Google Scholar 

  25. Jain, R.: Normalization of tumor vasculature: an emerging concept in antiangiogenic therapy. Science 307(5706), 58 (2005)

    Article  Google Scholar 

  26. Baxter, L., Jain, R.: Transport of fluid and macromolecules in tumors. I. Role of interstitial pressure and convection. Microvasc. Res. 37(1), 77 (1989)

    Article  Google Scholar 

  27. Baxter, L., Jain, R.: Transport of fluid and macromolecules in tumors. II. Role of heterogeneous perfusion and lymphatics. Microvasc. Res. 40(2), 246 (1990)

    Article  Google Scholar 

  28. Urcun, S., Rohan, P.Y., Skalli, W., Nassoy, P., Bordas, S.P., Sciumè, G.: Quantifying the role of mechanics in the free and encapsulated growth of cancer spheroids, bioRxiv (2020). https://doi.org/10.1101/2020.06.09.142927. https://www.biorxiv.org/content/early/2020/06/11/2020.06.09.142927

  29. Alessandri, K., Sarangi, B.R., Gurchenkov, V.V., Sinha, B., Kießling, T.R., Fetler, L., Rico, F., Scheuring, S., Lamaze, C., Simon, A., Geraldo, S., Vignjević, D., Doméjean, H., Rolland, L., Funfak, A., Bibette, J., Bremond, N., Nassoy, P.: Cellular capsules as a tool for multicellular spheroid production and for investigating the mechanics of tumor progression in vitro. Proc. Natl. Acad. Sci. 110(37), 14843 (2013)

    Article  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Giuseppe Sciumè.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendix: Chemotactic-Fickian model for EC diffusive velocity

Appendix: Chemotactic-Fickian model for EC diffusive velocity

Let us assume that all non-endothelial cell species (e.g., non-endothelial host cells, chemical species, etc.) can be modeled jointly as a species named H, the dominant species in h. Consequently let us treat species E as a diluted species in h. For such a situation, neglecting body forces potential and assuming isothermal condition, TCAT [11] provides from simplified entropy inequality the following force–flux pair:

$$\begin{aligned} \underbrace{\frac{1}{\theta }{\rho ^h}{\varepsilon ^h}{\omega ^{\overline{Eh} }}{{\mathbf{u}}^{\overline{\overline{Eh}} }}}_{{\mathrm{flux}}} \cdot \underbrace{\nabla \left( {{\mu ^{\overline{Eh} }} - {\mu ^{\overline{Hh} }}} \right) }_{{\mathrm{force}}} \ge 0, \end{aligned}$$
(91)

where \(\theta\) is the temperature and \(\nabla {\mu ^{\overline{Eh}}}\) and \(\nabla {\mu ^{\overline{Hh}}}\) are the chemical potentials of endothelial and non-endothelial species, respectively. At equilibrium, both force and flux terms are zero. Conversely, near equilibrium a first-order closure approximation for the force–flux relationship reads

$$\begin{aligned} {\omega ^{\overline{Eh} }}{{\mathbf{u}}^{\overline{\overline{Eh}} }} = - {\chi ^{\overline{\overline{Eh}} }}{\chi ^{\overline{\overline{Hh}} }}{{\mathbf{D}}^{Eh}} \cdot \nabla \left( {{\mu ^{\overline{Eh} }} - {\mu ^{\overline{Hh} }}} \right), \end{aligned}$$
(92)

where \({{\mathbf{D}}^{Eh}}\) is a second-order symmetric tensor, and \({\chi ^{\overline{\overline{ih}}}}\) is the molar fraction of species i. The Gibbs–Duhem equation for this binary mixture reads

$$\begin{aligned} \eta \overline{\overline{^h}} \nabla \theta \overline{\overline{^h}} - \nabla {p^h} + \sum \nolimits _i {{\rho ^h}{\omega ^{\overline{ih} }}\nabla {\mu ^{\overline{ih} }}} = {\mathbf{0}}. \end{aligned}$$
(93)

The expected pressure gradient in the phase h is relatively weak. Hence, for the isothermal case considered, Eq. (93) reduces for a binary mixture to

$$\begin{aligned} {\rho ^h}{\omega ^{\overline{Eh} }}\nabla {\mu ^{\overline{Eh} }} + {\rho ^h}{\omega ^{\overline{Hh} }}\nabla {\mu ^{\overline{Hh} }} = {\mathbf{0}}. \end{aligned}$$
(94)

Equation (94) allows us to obtain the gradient of the chemical potential of species H as

$$\begin{aligned} \nabla {\mu ^{\overline{Hh} }} = - \frac{{{\omega ^{\overline{Eh} }}}}{{{\omega ^{\overline{Hh} }}}}\nabla {\mu ^{\overline{Eh} }}. \end{aligned}$$
(95)

With \({\omega ^{\overline{Eh} }} \ll {\omega ^{\overline{Hh} }}\) it follows that \(\nabla {\mu ^{\overline{Hh} }} \ll \nabla {\mu ^{\overline{Eh} }}\) and consequently Eq. (92) reduces to

$$\begin{aligned} {\omega ^{\overline{Eh} }}{{\mathbf{u}}^{\overline{\overline{Eh}} }} = - {\chi ^{\overline{\overline{Eh}} }}{\chi ^{\overline{\overline{Hh}} }}{{\mathbf{D}}^{Eh}} \cdot \nabla {\mu ^{\overline{Eh} }}. \end{aligned}$$
(96)

To gain usefulness of the previous equation, a relationship between the macroscale chemical potential of species E and its mass fraction is needed. The macroscale chemical potential for the species E can be written as

$$\begin{aligned} {\mu ^{\overline{Eh} }} = \mu _0^{\overline{Eh} }\left( {{p^h},\theta } \right) + \frac{{R\theta }}{{{M_E}}}\ln \left( {{\chi ^{\overline{\overline{Eh}} }}{\gamma ^{\overline{\overline{Eh}} }}} \right), \end{aligned}$$
(97)

where \(\mu _0^{\overline{Eh} }\left( {{p^h},\theta } \right)\) is a reference chemical potential for species E, R is the ideal gas constant, \(M_E\) is the molar mass of species E, and \({\gamma ^{\overline{\overline{Eh}} }}\) is the macroscale activity coefficient. With the system in isothermal condition (\(\theta = {\theta _0}\)) and the impact of pressure gradient of phase h assumed negligible, differentiating this expression in space gives

$$\begin{aligned} \nabla {\mu ^{\overline{Eh} }} = \frac{{R{\theta _0}}}{{{M_E}}}\frac{1}{{{\chi ^{\overline{\overline{Eh}} }}}}\nabla {\chi ^{\overline{\overline{Eh}} }} + \frac{{R{\theta _0}}}{{{M_E}}}\frac{1}{{{\gamma ^{\overline{\overline{Eh}} }}}}\nabla {\gamma ^{\overline{\overline{Eh}} }}. \end{aligned}$$
(98)

For dilute species, the macroscale activity coefficient is usually assumed constant and equal to 1. To account for chemotaxis we set here an activity coefficient linearly dependent on the mass fraction of TAF in h. We assume here local chemical equilibrium, so despite the mass fraction of TAF in h, \({\omega ^{\overline{Ah}}}\), not being a primary variable of the model, its value can be linearly related (as a first approximation) to the mass fraction of TAF in the adjacent IF phase, \({\omega ^{\overline{Ah}}}\propto {\omega ^{\overline{Al}}}\). This allows us to assume that the activity coefficient, \({\gamma ^{\overline{\overline{Eh}} }}\), is linearly dependent on the mass fraction of TAF in l. Thanks to short-range diffusion and molecular signaling the TAF in the phase l interferes (via the hl interface) with endothelial cells modifying their activity coefficient. The following relationship is assumed, with c the constant chemotactic coefficient and \({\chi ^{\overline{\overline{Al}}}}\) the molar fraction of TAF in IF:

$$\begin{aligned} {\gamma ^{\overline{\overline{Eh}} }} = 1 - c{\chi ^{\overline{\overline{Al}} }}. \end{aligned}$$
(99)

Introducing Eq. (99) into Eq. (98) gives

$$\begin{aligned} \nabla {\mu ^{\overline{Eh} }} = \frac{{R{\theta _0}}}{{{M_E}}}\frac{1}{{{\chi ^{\overline{\overline{Eh}} }}}}\nabla {\chi ^{\overline{\overline{Eh}} }} - \frac{{R{\theta _0}}}{{{M_E}}}\frac{c}{{\left( {1 - c{\chi ^{\overline{\overline{Al}} }}} \right) }}\nabla {\chi ^{\overline{\overline{Al}} }}. \end{aligned}$$
(100)

We reasonably assume here that the molar masses of phases h and l are weakly affected by variation of species concentration. This allows us to assume constant molar masses of phases h and l and to express the molar fraction of the species E, A and H as functions of the respective mass fractions:

$$\begin{aligned} {\chi ^{\overline{\overline{Eh}} }} = \frac{{{M_h}}}{{{M_E}}}{\omega ^{\overline{Eh} }}\mathop {}\nolimits _{}^{} \mathop {}\nolimits _{}^{} \mathop {}\nolimits _{}^{} \mathop {}\nolimits _{}^{}\!\!\!\!\!,\; {\chi ^{\overline{\overline{Al}} }} = \frac{{{M_l}}}{{{M_A}}}{\omega ^{\overline{Al} }}\mathop {}\nolimits _{}^{} \mathop {}\nolimits _{}^{} \mathop {}\nolimits _{}^{} \mathop {}\nolimits _{}^{}\!\!\!\!\!,\, {\chi ^{\overline{\overline{Hh}} }} = \frac{{{M_h}}}{{{M_H}}}{\omega ^{\overline{Hh} }}. \end{aligned}$$
(101)

Introducing the first two relationships of Eq. (101) into Eq. (100) and setting \(C = c\frac{{{M_l}}}{{{M_A}}}\) gives

$$\begin{aligned} \nabla {\mu ^{\overline{Eh} }} = \frac{{R{\theta _0}}}{{{M_E}}}\frac{1}{{{\omega ^{\overline{Eh} }}}}\nabla {\omega ^{\overline{Eh} }} - \frac{{R{\theta _0}}}{{{M_E}}}\frac{C}{{\left( {1 + C{\omega ^{\overline{Al} }}} \right) }}\nabla {\omega ^{\overline{Al} }}. \end{aligned}$$
(102)

We now introduce Eq. (102) into Eq. (96) and express \({\chi ^{\overline{\overline{i\alpha }} }}\) as function of \({\omega ^{\overline{i\alpha } }}\). After some calculations, we obtain

$$\begin{aligned} {\omega ^{\overline{Eh} }}{{\mathbf{u}}^{\overline{\overline{Eh}} }} = - \underbrace{\frac{{{{\left( {{M_h}} \right) }^2}}}{{{M_E}{M_H}}}{\omega ^{\overline{Hh} }}\frac{{R{\theta _0}}}{{{M_E}}}{{\mathbf{D}}^{Eh}}}_{ \cong {{\text { constant second{-}order tensor}}}} \cdot \nabla {\omega ^{\overline{Eh} }} + \frac{{C{\omega ^{\overline{Eh} }}}}{{\left( {1 + C{\omega ^{\overline{Al} }}} \right) }}\underbrace{\frac{{{{\left( {{M_h}} \right) }^2}}}{{{M_E}{M_H}}}{\omega ^{\overline{Hh} }}\frac{{R{\theta _0}}}{{{M_E}}}{{\mathbf{D}}^{Eh}}}_{ \cong {{\text { constant second{-}order tensor}}}} \cdot \nabla {\omega ^{\overline{Al} }}. \end{aligned}$$
(103)

As shown in the previous equation, some quantities are expected to stay almost constant, resulting in always \({\omega ^{\overline{Hh} }} \cong 1\). This observation allows us to rewrite the previous equation in a simplified form

$$\begin{aligned} {\omega ^{\overline{Eh} }}{{\mathbf{u}}^{\overline{\overline{Eh}} }} = - {{{\hat{\mathbf{D}}}}^{Eh}} \cdot \nabla {\omega ^{\overline{Eh} }} + \frac{{C{\omega ^{\overline{Eh} }}}}{{\left( {1 + C{\omega ^{\overline{Al} }}} \right) }}{{{\hat{\mathbf{D}}}}^{Eh}} \cdot \nabla {\omega ^{\overline{Al} }}, \end{aligned}$$
(104)

where the diffusivity tensor \({{{\hat{\mathbf{D}}}}^{Eh}}\) reads

$$\begin{aligned} {{{\hat{\mathbf{D}}}}^{Eh}} = {\left( {\frac{{{M_h}}}{{{M_E}}}} \right) ^2}\frac{{R{\theta _0}}}{{{M_H}}}{\omega ^{\overline{Hh} }}{{\mathbf{D}}^{Eh}}. \end{aligned}$$
(105)

We assume here an isotropic effective diffusivity which linearly increases with the volume fraction of phase h. Therefore, Eq. (104) can be rewritten in the form

$$\begin{aligned} {\omega ^{\overline{Eh} }}{{\mathbf{u}}^{\overline{\overline{Eh}} }} = - D_{\mathrm{{eff}}}^{\overline{Eh} } \cdot \nabla {\omega ^{\overline{Eh} }} + \frac{{C{\omega ^{\overline{Eh} }}}}{{\left( {1 + C{\omega ^{\overline{Al} }}} \right) }}D_{\mathrm{{eff}}}^{\overline{Eh} } \cdot \nabla {\omega ^{\overline{Al} }}, \end{aligned}$$
(106)

where indicating with \(D_0^{\overline{Eh}}\) the bulk diffusivity of endothelial cells in h, the effective diffusivity reads

$$\begin{aligned} D_{\mathrm{{eff}}}^{\overline{Eh} } = D_0^{\overline{Eh} }{\varepsilon ^h}. \end{aligned}$$
(107)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Sciumè, G. Mechanistic modeling of vascular tumor growth: an extension of Biot’s theory to hierarchical bi-compartment porous medium systems. Acta Mech 232, 1445–1478 (2021). https://doi.org/10.1007/s00707-020-02908-z

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00707-020-02908-z

Navigation