Skip to main content
Log in

The interplay between tissue growth and scaffold degradation in engineered tissue constructs

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

Abstract

In vitro tissue engineering is emerging as a potential tool to meet the high demand for replacement tissue, caused by the increased incidence of tissue degeneration and damage. A key challenge in this field is ensuring that the mechanical properties of the engineered tissue are appropriate for the in vivo environment. Achieving this goal will require detailed understanding of the interplay between cell proliferation, extracellular matrix (ECM) deposition and scaffold degradation. In this paper, we use a mathematical model (based upon a multiphase continuum framework) to investigate the interplay between tissue growth and scaffold degradation during tissue construct evolution in vitro. Our model accommodates a cell population and culture medium, modelled as viscous fluids, together with a porous scaffold and ECM deposited by the cells, represented as rigid porous materials. We focus on tissue growth within a perfusion bioreactor system, and investigate how the predicted tissue composition is altered under the influence of (1) differential interactions between cells and the supporting scaffold and their associated ECM, (2) scaffold degradation, and (3) mechanotransduction-regulated cell proliferation and ECM deposition. Numerical simulation of the model equations reveals that scaffold heterogeneity typical of that obtained from \(\mu \)CT scans of tissue engineering scaffolds can lead to significant variation in the flow-induced mechanical stimuli experienced by cells seeded in the scaffold. This leads to strong heterogeneity in the deposition of ECM. Furthermore, preferential adherence of cells to the ECM in favour of the artificial scaffold appears to have no significant influence on the eventual construct composition; adherence of cells to these supporting structures does, however, lead to cell and ECM distributions which mimic and exaggerate the heterogeneity of the underlying scaffold. Such phenomena have important ramifications for the mechanical integrity of engineered tissue constructs and their suitability for implantation in vivo.

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.

Institutional subscriptions

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8

Similar content being viewed by others

References

  • Adachi T, Osako Y, Tanaka M, Hojo M, Hollister SJ (2006) Framework for optimal design of porous scaffold microstructure by compututational simulation of bone regeneration. Biomaterials 27:3964–3972

    Article  Google Scholar 

  • Ahearne M, Wilson SL, Liu K-K, Rauz S, El Haj AJ, Yang Y (2010) Influence of cell and collagen concentration on the cell-matrix mechanical relationship in a corneal stroma wound healing model. Exp Eye Res 91:584–591

    Article  Google Scholar 

  • Araujo RP, McElwain DLS (2005) A mixture theory for the genesis of residual stresses in growing tissues i: a general formulation. SIAM J Appl Math 65(4):1261–1284

    Article  MathSciNet  MATH  Google Scholar 

  • Bakker A, Klein-Nulend J, Burger E (2004a) Shear stress inhibits while disuse promotes osteocyte apoptosis. Biochem Biophys Res Commun 320(4):1163–1168

    Article  Google Scholar 

  • Bakker A, Klein-Nulend J, Burger E (2004b) Shear stress inhibits while disuse promotes osteocyte apoptosis. Biochem Biophys Res Comm 320:1163–1168

    Article  Google Scholar 

  • Breward CJW, Byrne HM, Lewis CE (2002) The role of cell–cell interactions in a two-phase model for avascular tumour growth. J Math Biol 45:125–152

    Article  MathSciNet  MATH  Google Scholar 

  • Burdick JA, Mauck RL (eds) (2010) Biomaterials for tissue engineering applications: a review of the past and future trends. Springer, Wien

  • Byrne DP, Lacroix D, Planell JA, Kelly DJ, Prendergast PJ (2007) Simulation of tissue differentiation in a scaffold as a function of porosity, Young’s modulus and dissolution rate: application of mechanobiological models in tissue engineering. Biomaterials 28(36):5544–5554

    Article  Google Scholar 

  • Byrne HM, Preziosi L (2003) Modelling solid tumour growth using the theory of mixtures. Math Med Biol 20(4):341–366

    Article  MathSciNet  MATH  Google Scholar 

  • Cartmell SH, El Haj AJ (2005) Mechanical bioreactors for tissue engineering. In: Chaudhuri J, Al-Rubeai M (eds) Bioreactors for tissue engineering: principles, design and operation, chapt. 8. Springer, Dordrecht, pp 193–209

    Chapter  Google Scholar 

  • Chaplain MAJ, Graziano L, Preziosi L (2006) Mathematical modelling of the loss of tissue compression responsiveness and its role in solid tumour development. Math Med Biol 23(3):197

    Article  MATH  Google Scholar 

  • Cowin SC (2000) How is a tissue built? J Biomech Eng 122:553

    Article  Google Scholar 

  • Cowin SC (2004) Tissue growth and remodeling. Ann Rev Biomed Eng 6(1):77–107

    Article  Google Scholar 

  • Curtis A, Riehle M (2001) Tissue engineering: the biophysical background. Phys Med Biol 46:47–65

    Article  Google Scholar 

  • Drew DA, Segel LA (1971) Averaged equations for two-phase flows. Studies Appl Math 50:205–231

    MATH  Google Scholar 

  • El Haj AJ, Minter SL, Rawlinson SC, Suswillo R, Lanyon LE (1990) Cellular responses to mechanical loading in vitro. J Bone Min Res 5(9):923–932

    Article  Google Scholar 

  • Franks SJ, King JR (2003) Interactions between a uniformly proliferating tumour and its surroundings: uniform material properties. Math Med Biol 20:47–89

    Article  MATH  Google Scholar 

  • Freed LE, Vunjak-Novakovic G (1998) Culture of organized cell communities. Adv Drug Deliv Rev 33:15–30

    Article  Google Scholar 

  • Freed LE, Vunjak-Novakovic G, Biron RJ, Eagles DB, Lesnoy DC, Barlow SK, Langer R (1994) Biodegradable polymer scaffolds for tissue engineering. Nat Biotech 12(7):689–693

    Article  Google Scholar 

  • Fung YC (1991) What are residual stresses doing in our blood vesels? Ann Biomed Eng 19:237–249

    Article  MathSciNet  Google Scholar 

  • Haider MA, Olander JE, Arnold RF, Marous DR, McLamb AJ, Thompson KC, Woodruff WR, Haugh JM (2010) A phenomenological mixture model for biosynthesis and linking of cartilage extracellular matrix in scaffolds seeded with chondrocytes. Biomech Model Mechanobiol 1–10. ISSN 1617–7959

  • Han Y, Cowin SC, Schaffler MB, Weinbaaum S (2004) Mechanotransduction and strain amplification in osteocyte cell processes. Proc Natl Acad Sci 101(47):16689–16694

    Article  Google Scholar 

  • Holzapfel GA, Ogden RW (2006) Mechanics of biological tissue. Springer, Berlin

    Book  Google Scholar 

  • Kelly DJ, Prendergast PJ (2003) Effect of a degraded core on the mechanical behaviour of tissue-engineered cartilage constructs: a poro-elastic finite element analysis. Med Biol Eng Comp 42:9–13

    Article  Google Scholar 

  • Klein-Nulend J, Roelofsen J, Sterck JG, Semeins CM, Burger EH (1995) Mechanical loading stimulates the release of transforming growth factor-beta activity by cultured mouse calvariae and periosteal cells. J Cell Physiol 163(1):115–119

    Article  Google Scholar 

  • Wu L, Ding J (2004) In vitro degradation of three-dimensional porous poly(d, l-lactide-co-glycolide) scaffolds for tissue engineering. Biomaterials 25(27):5821–5830

    Article  Google Scholar 

  • Landman KA, Please CP (2001) Tumour dynamics and necrosis: surface tension and stability. IMA J Math Appl Med Biol 18(2):131–158

    Article  MATH  Google Scholar 

  • Lemon G, King JR (2007) Multiphase modelling of cell behaviour on artificial scaffolds: effects of nutrient depletion and spatially nonuniform porosity. Math Med Biol 24(1):57

    Article  MATH  Google Scholar 

  • Lemon G, King JR, Byrne HM, Jensen OE, Shakesheff K (2006) Multiphase modelling of tissue growth using the theory of mixtures. J Math Biol 52(2):571–594

    Article  MathSciNet  MATH  Google Scholar 

  • Lewis MC, Macarthur BD, Malda J, Pettet G, Please CP (2005) Heterogeneous proliferation within engineered cartilaginous tissue: the role of oxygen tension. Biotech Bioeng 91(5):607–615

    Article  Google Scholar 

  • Lubkin SR, Jackson T (2002) Multiphase mechanics of capsule formation in tumors. J Biomech Eng 124:237

    Article  Google Scholar 

  • Martin I, Wendt D, Heberer M (2004) The role of bioreactors in tissue engineering. Trends Biotechnol 22(2):80–86

    Article  Google Scholar 

  • Nikolovski J, Mooney DJ (2000) Smooth muscle cell adhesion to tissue engineering scaffolds. Biomaterials 21(20):2025–2032

    Article  Google Scholar 

  • O’Dea RD, Waters SL, Byrne HM (2008) A two-fluid model for tissue growth within a dynamic flow environment. Eur J Appl Math 19(641):607–634

    Article  MathSciNet  MATH  Google Scholar 

  • O’Dea RD, Waters SL, Byrne HM (2010) A three phase model for tissue construct growth in a perfusion bioreactor. J Math Med Biol 27(2):95–127

    Article  MathSciNet  MATH  Google Scholar 

  • O’Dea RD, Waters SL, Byrne HM (2012) Modelling tissue growth in bioreactors: a review. In: Geris L (ed) Computational modeling in tissue engineering. Springer, Berlin

    Google Scholar 

  • Osborne JM, O’Dea RD, Whiteley JP, Byrne HM, Waters SL (2010) The influence of bioreactor geometry and the mechanical environment on engineered tissues. J Biomech Eng 132(5), doi:10.1115/1.4001160

  • Osborne JM, Whiteley JP (2010) A numerical method for the multiphase viscous flow equations. Comput Methods Appl Mech Engrg 199:3402–3417. doi:10.1016/j.cma.2010.07.011

    Article  MathSciNet  MATH  Google Scholar 

  • Preziosi L, Tosin A (2009) Multiphase and multiscale trends in cancer modelling. Math Model Nat Phenom 4(3):1–11, ISSN 0973–5348

    Google Scholar 

  • Roelofsen J, Klein-Nulend J, Burger EH (1995) Mechanical stimulation by intermittent hydrostatic compression promotes bone-specific gene expression in vitro. J Biomech 28(12):1493–1503

    Article  Google Scholar 

  • Roose T, Neti PA, Munn LL, Boucher Y, Jain RK (2003) Solid stress generated by spheroid growth estimated using a poroelasticity model. Microvasc Res 66:204–212

    Article  Google Scholar 

  • Sanz-Herrera JA, García-Aznar JM, Doblaré M (2008) On scaffold designing for bone regeneration: a computational multiscale approach. Acta Biomat (online preprint)

  • Sipe JD (2002) Tissue engineering and reparative medicine. Ann N Y Acad Sci 961:1–9

    Article  Google Scholar 

  • Skalak R, Zargaryan S, Jain RK, Netti PA, Hoger A (1996) Compatibility and the genesis of residual stress by volumetric growth. J Math Biol 34:889–914

    MATH  Google Scholar 

  • Urban JPG (1994) The chondrocyte: a cell under pressure. Rheumatology 33(10):901–908

    Article  Google Scholar 

  • VonNeumann J, Richtmyer RD (1950) A method for the numerical calculation of hydrodynamic shocks. J Appl Phys 21:232

    Article  MathSciNet  Google Scholar 

  • Wang QG, Nguyen B, Thomas CR, Zhang Z, El Haj AJ, Kuiper NJ (2010) Molecular profiling of single cells in response to mechanical force: comparison of chondrocytes, chondrons and encapsulated chondrocytes. Biomaterials 31(7):1619–1625

    Article  Google Scholar 

  • Wilson DJ, King JR, Byrne HM (2007) Modelling scaffold occupation by a growing, nutrient-rich tissue. Math Models Methods Appl Sci 17:1721

    Article  MathSciNet  MATH  Google Scholar 

  • Yang Y, El Haj AJ (2006) Biodegradable scaffolds—delivery systems for cell therapies. Expert Opin Biol Ther 6(5):485–498

    Article  Google Scholar 

  • You J, Yellowley CE, Donahue HJ, Zhang Y, Chen Q, Jacobs CR (2000) Substrate deformation levels associated with routine physical activity are less stimulatory to bone cells relative to loading-induced oscillatory fluid flow. J Biomech Eng 122:377–393

    Article  Google Scholar 

  • You L, Cowin SC, Schaffler MB, Weinbaum S (2001) A model for strain amplification in the actin cytoskeleton of osteocytes due to fluid drag on pericellular matrix. J Biomech 34(11):1375–1386

    Article  Google Scholar 

  • Yourek G, Al-Hadlaq A, Patel R, McCormick S, Reilly GC, Mao JJ (2004) Nanophysical properties of living cells. In: Stroscio Michael A, Dutta Mitra, He Bin (eds) Biological nanostructures and applications of nanostructures in biology, bioelectric engineering. Springer, US, pp 69–97

    Chapter  Google Scholar 

Download references

Acknowledgments

This work was supported by funding from the EPSRC in the form of a Ph.D studentship (RDO) and an Advanced Research Fellowship (SLW). It was also supported by the EPSRC/BBSRC-funded OCISB project BB/D020190/1 (JMO), and based on work supported in part by Award KUK-013-04 made by King Abdullah University of Science and Technology (HMB). We are also grateful to E. Baas, ISTM, Keele University for the provision of experimental data.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to R. D. O’Dea.

Additional information

R. D. O’Dea and J. M. Osborne are the ioint first authors.

Appendix: Model derivation

Appendix: Model derivation

We consider a bioreactor of length \(L^*\) and width \(h^*\), modelled as a two-dimensional channel containing a mixture of four interacting phases, representing cells, culture medium, PLLA scaffold and ECM and denote these via a subscript \(i=n,\ w,\ s,\ e\), respectively. The viscosity of any fluid phase is denoted \(\mu ^*_i\), and the typical timescale for tissue growth (comprising both cell proliferation and ECM deposition) is denoted \(K^*\). Asterisks distinguish dimensional quantities from their dimensionless equivalents.

We introduce a Cartesian coordinate system \(L^*\varvec{x}=L^*(x,y)\) and time \(K^*t\) and the channel occupies the dimensionless region \(0\le x \le 1,\,0\le y\le h=h^*/L^*\). The volume fraction of each phase is denoted \(\theta _i\), while the dimensionless volume-averaged velocities, pressures and stress tensors of the each phase are denoted \(K^*L^*\varvec{u}_i=K^*L^*(u_i,v_i),\,K^*\mu _w^*p_i\) and \(K^*\mu _w^*\varvec{\sigma }^i\). Tissue growth, scaffold degradation and ECM deposition are captured via material transfer functions \(K^*S_i\). We assume that all dimensionless dependent variables are functions of \(\varvec{x}\) and \(t\).

The model is constructed by considering mass and momentum balances for each phase, assuming that each phase is incompressible, with equal density, and neglecting inertial effects; the equations governing the \(i\mathrm{th }\) phase (with volume fraction \(\theta _i\)) are as follows [see Lemon et al. (2006), O’Dea et al. (2010), Osborne et al. (2010)]:

$$\begin{aligned} \frac{\partial \theta _{i}}{\partial t}+\nabla \cdot (\theta _{i} \varvec{u}_{i})&= S_{i}(\theta _{k},p_{k},\varvec{u}_{k})\,, \end{aligned}$$
(21)
$$\begin{aligned} \nabla \cdot \left(\theta _{i} \varvec{\sigma }^{i}\right) + \sum _{j\ne i} \mathbf{F}^{ij}&= \varvec{0}. \end{aligned}$$
(22)

Additional conservation conditions may be obtained by summing over all phases and exploiting the no-voids condition \(\sum _i \theta _i=1\).

In Eq. (21) \(K^* S_i\) is the net material production term associated with phase \(i\) (mass conservation demands that \(\sum S_i=0\)); in (22), \(K^*\mu _w^*/L^* \mathbf F ^{ij}\) is the interphase force exerted by phase \(j\) on phase \(i\), obeying \(\mathbf F ^{ij}=-\mathbf F ^{ji}\). These interphase forces comprise interphase viscous drag (with drag coefficient \(\mu _w^*/L^{*2} k\)) and active forces, the latter being embodied within extra pressures which arise due to cell–cell, cell–ECM and cell–scaffold interactions; interactions between the culture medium and scaffold phases are assumed to involve only viscous drag. The mechanics of this four phase formulation is simplified by lumping the scaffold and ECM components into a single ‘substrate’ phase, denoted \(\theta _S=\theta _s+\theta _e\), and modelled as a rigid porous material. For notational convenience, in this Appendix, we employ the subscript \(S\) to denote the substrate, in preference to \(\Theta \). Separate mass conservation equations are nevertheless employed for \(\theta _s\) and \(\theta _e\) to track their individual evolution.

The cell population and culture medium are represented as distinct viscous fluids, modelled by standard viscous stress tensors; the rigidity of the substrate implies \(\varvec{u}_S \!=\!\varvec{0}\). These constitutive assumptions are embodied in the following equations.

$$\begin{aligned} \varvec{\sigma }^i&= -p_i\mathbf I + \mu _i\left(\nabla \varvec{u}_i+\nabla \varvec{u}_i^T\right)-\tfrac{2}{3}\left(\nabla \cdot \varvec{u}_i\right)\mathbf I \,, \quad \text{ for} \quad i=n,w\,, \end{aligned}$$
(23)
$$\begin{aligned} \mathbf F ^{ij}&= \left(p_{w}\!+\!\psi _{ij}\right)\left(\theta _{j}\nabla \theta _{i}\!-\!\theta _{i} \nabla \theta _{j}\right)\!+\!k\theta _{i}\theta _{j}(\varvec{u}_{j}\!-\!\varvec{u}_{i})\,, \quad \text{ for} \quad i,j\!=\!n,w,S\,,\end{aligned}$$
(24)
$$\begin{aligned} p_n&= p_w + \theta ^2_n\Sigma _n + \theta _n\psi _{nS}\,, \end{aligned}$$
(25)

wherein \(\mu _i\) are the dimensionless viscosities of each phase, and \(\Sigma _{n}\) and \(\psi _{nS}\) are defined

$$\begin{aligned} \Sigma _{n} = - \nu + \frac{\delta _{a}\theta _{n}}{\theta _{w}}\,\quad \text{ and} \quad \psi _{nS} = -\chi + \frac{\delta _{b}\theta _{n}}{\theta _{w}}. \end{aligned}$$
(26)

In Eq. (26) \(\nu ,\ \chi ,\ \delta _a,\ \delta _b>0\) dictate the cells’ tendency to aggregate, their affinity for the scaffold/ECM and the strength of cell–cell/cell–scaffold repulsion. In a more general formulation, the coefficient of viscous drag \(k\)  between two phases \(i\)  and \(j\)  varies depending upon the phases under consideration and may depend upon their respective volume fractions or other state variables. A suitable representation is to replace \(k\)  by, say, \(k_{ij}(\theta _i,\theta _j)\)  (obeying \(k_{ij}=k_{ji}\)). Full details and discussion of the above choice of interphase interaction terms may be found in Lemon et al. (2006).

Figure 9 depicts the two-dimensional model of the bioreactor, together with appropriate boundary conditions. These correspond to no-slip and no-penetration of cells or culture medium through the channel walls, a pressure-driven flow imposed via up- and downstream pressures \(K^*\mu _w^* P_U\) and \(K^*\mu _w^* P_D\), partitioned normal stress conditions and fully-developed flow at \(x=0,1\).

Fig. 9
figure 9

The two-dimensional domain, outward-pointing normal \(\hat{\varvec{n}}\) and associated boundary conditions. The arrows indicate the perfusion direction in the case of dimensionless up- and downstream pressures \(P_U\) and \(P_D\) obeying \(P_U>P_D\)

We now simplify the two-dimensional equations by considering the limit for which the aspect ratio of the bioreactor is asymptotically small \((h\ll 1)\). We remark that, since the culture medium volume fraction may be eliminated via \(\theta _w=1-\theta _n-\theta _s-\theta _e\) and the substrate is rigid, we need consider momentum conservation equations for the fluid (cell and culture medium) phases, only.

Following O’Dea et al. (2010), the reduced model is obtained by rescaling according to:

$$\begin{aligned} y=h\overline{y}, \quad v_i=h \overline{v}_i,\quad p_i=\overline{p}_i/h^2, \end{aligned}$$
(27)

and averaging across the channel in the transverse direction (imposing the boundary conditions at \(y=0,h\) depicted in Fig. 9). We find that the pressure and the volume fraction of each phase are functions of \(x\) and \(t\) only and the flow of cells and culture medium is unidirectional at leading order \((\overline{v}_i=0)\). Expressions for the averaged axial velocities \(\langle \overline{u}_w \rangle \) and \(\langle \overline{u}_n \rangle \) are obtained from the remaining momentum equations, on substitution of which into the (averaged) mass conservation equations (dropping the overbars), we obtain the following system of coupled partial differential equations for the volume fractions \(\theta _e(x,t),\,\theta _s(x,t),\ \theta _n(x,t)\) and the culture medium pressure, \(p_w(x,t)\):

$$\begin{aligned} \frac{\partial \theta _s}{\partial t}&= S_s\, ,\end{aligned}$$
(28)
$$\begin{aligned} \frac{\partial \theta _e}{\partial t}&= S_e\, ,\end{aligned}$$
(29)
$$\begin{aligned} \frac{\partial \theta _{n}}{\partial t}+\frac{1}{12}\frac{\partial }{\partial x} \left((1-\theta _s-\theta _e-\theta _{n}) \frac{\partial p_w}{\partial x} \right)&= S_{n} \,, \end{aligned}$$
(30)
$$\begin{aligned}&\frac{\partial }{\partial x} \left\{ \left(\theta _{n}+ \mu _n(1-\theta _s-\theta _e-\theta _{n})\right)\frac{\partial p_w}{\partial x}\right\} \nonumber \\& \quad \quad \,\,\, + \frac{\partial }{\partial x} \left\{ \frac{\partial (\theta _{n}^2 {\Sigma _n})}{\partial x}+ 2\theta _{n}\psi _{nS}\frac{\partial (\theta _e+\theta _s)}{\partial x} + \theta _{n}(\theta _e+\theta _s)\frac{\partial \psi _{nS}}{\partial x}\right\} = 0\,, \end{aligned}$$
(31)

in which \(\mu _n\) is the relative viscosity of the cell and culture medium phases. The extra pressures \(\Sigma _n\) and \(\psi _{nS}\) are scaled according to Eq.  (27) so that these interactions are retained at leading order, which implies \((\nu ,\delta _a,\chi ,\delta _b)=(\bar{\nu },\bar{\delta }_a,\bar{\chi },\bar{\delta }_b)/h^2\); the remaining parameters are \(\mathcal O (1)\). Equations (28)–(31) embody conservation of mass for the ECM, PLLA scaffold and cell phases, and the multiphase mixture. Dropping the subscripts on the interaction functions \(\Sigma _n\)  and \(\psi _{nS}\)  gives the equations stated in the main text.

Under the rescaling, (27), the boundary conditions shown in Fig. 9 become:

$$\begin{aligned} p_{w}&= P_{U} \quad \text{ at} \quad x =0 \,,\end{aligned}$$
(32)
$$\begin{aligned} p_{w}&= P_{D} \quad \text{ at} \quad x =1 \,,\end{aligned}$$
(33)
$$\begin{aligned} \frac{\partial \theta _{n}}{\partial x}&= 0 \quad \quad \text{ at} \quad x =0,1. \end{aligned}$$
(34)

Rights and permissions

Reprints and permissions

About this article

Cite this article

O’Dea, R.D., Osborne, J.M., El Haj, A.J. et al. The interplay between tissue growth and scaffold degradation in engineered tissue constructs. J. Math. Biol. 67, 1199–1225 (2013). https://doi.org/10.1007/s00285-012-0587-9

Download citation

  • Received:

  • Revised:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00285-012-0587-9

Keywords

Mathematics Subject Classification (2000)

Navigation