Skip to main content
Log in

The Impact of Cell Density and Mutations in a Model of Multidrug Resistance in Solid Tumors

  • Original Article
  • Published:
Bulletin of Mathematical Biology Aims and scope Submit manuscript

Abstract

In this paper we develop a mathematical framework for describing multidrug resistance in cancer. To reflect the complexity of the underlying interplay between cancer cells and the therapeutic agent, we assume that the resistance level is a continuous parameter. Our model is written as a system of integro-differential equations that are parameterized by the resistance level. This model incorporates the cell density and mutation dependence. Analysis and simulations of the model demonstrate how the dynamics evolves to a selection of one or more traits corresponding to different levels of resistance. The emerging limit distribution with nonzero variance is the desirable modeling outcome as it represents tumor heterogeneity.

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
Fig. 15

Similar content being viewed by others

References

  • Bellomo, N. (1997). Nonlinear models and problems in applied sciences from differential quadrature to generalized collocation methods. Math. Comput. Model., 26(4), 13–34.

    Article  MathSciNet  MATH  Google Scholar 

  • Birkhead, B. G., Rakin, E. M., Gallivan, S., Dones, L., & Rubens, R. D. (1987). A mathematical model of the development of drug resistance to cancer chemotherapy. Eur. J. Cancer Clin. Oncol., 23(9), 1421–1427.

    Article  Google Scholar 

  • Bozic, I., Allen, B., & Nowak, M. A. (2012). Dynamics of targeted cancer therapy. Trends Mol. Med., 18(6), 311–316.

    Article  Google Scholar 

  • Brimacombe, K. R., Hall, M. D., Auld, D. S., Inglese, J., Austin, C. P., Gottesman, M. M., & Fung, K. L. (2009). A dual-fluorescence high-throughput cell line system for probing multidrug resistance. Assay Drug Dev. Technol., 7(3), 233–249.

    Article  Google Scholar 

  • Calsina, À., & Cuadrado, S. (2000). A model for the adaptive dynamics of the maturation age. Ecol. Model., 133, 33–43.

    Article  Google Scholar 

  • Calsina, À., & Cuadrado, S. (2004). Small mutation rate and evolutionary stable strategies in infinite dimensional adaptive dynamics. J. Math. Biol., 48, 135–159.

    Article  MathSciNet  MATH  Google Scholar 

  • Champagnat, N., Ferrière, R., & Méléard, S. (2006). Unifying evolutionary dynamics: from individual stochastic processes to macroscopic models. Theor. Popul. Biol., 69(3), 297–321.

    Article  MATH  Google Scholar 

  • Coldman, A. J., & Goldie, J. H. (1985). Role of mathematical modeling in protocol formulation in cancer chemotherapy. Cancer Treat. Rep., 69(10), 1041–1048.

    Google Scholar 

  • Coldman, A. J., & Goldie, J. H. (1986). A stochastic model for the origin and treatment of tumors containing drug-resistant cells. Bull. Math. Biol., 48(3–4), 279–292.

    Article  MathSciNet  MATH  Google Scholar 

  • Desvillettes, L., Jabin, P. E., Mischler, S., & Raoul, G. (2008). On selection dynamics for continuous structured populations. Commun. Math. Sci., 6(3), 729–747.

    Article  MathSciNet  MATH  Google Scholar 

  • Diekmann, O., Jabin, P. E., Mischler, S., & Perthame, B. (2005). The dynamics of adaptation: an illuminating example and a Hamilton–Jacobi approach. Theor. Popul. Biol., 67(4), 257–271.

    Article  MATH  Google Scholar 

  • Fister, K. R., & Panetta, J. C. (2003). Optimal control applied to competing chemotherapeutic cell-kill strategies. SIAM J. Appl. Math., 63(6), 1954–1971.

    Article  MathSciNet  MATH  Google Scholar 

  • Fodal, V., Pierobon, M., Liotta, L., & Petricoin, E. (2011). Mechanisms of cell adaptation: when and how do cancer cells develop chemoresistance? Cancer J., 17(2), 89–95.

    Article  Google Scholar 

  • Forys, U., & Marciniak-Czochra, A. (2003). Logistic equations in tumor growth modeling. Int. J. Appl. Math. Comput. Sci., 13(3), 317–325.

    MathSciNet  MATH  Google Scholar 

  • Gillet, J. P., & Gottesman, M. M. (2010). Mechanisms of multidrug reistance in cancer. Methods Mol. Biol., 596, 47–76.

    Article  Google Scholar 

  • Goldie, J. H., & Coldman, A. J. (1979). A mathematical model for relating the drug sensitivity of tumors to their spontaneous mutation rate. Cancer Treat. Rep., 63(11–12), 1727–1733.

    Google Scholar 

  • Goldie, J. H., & Coldman, A. J. (1983a). A model for resistance of tumor cells to cancer chemotherapeutic agents. Math. Biosci., 65, 291–307.

    Article  MATH  Google Scholar 

  • Goldie, J. H., & Coldman, A. J. (1983b). Quantative model for multiple levels of drug resistance in clinical tumors. Cancer Treat. Rep., 67(10), 923–931.

    Google Scholar 

  • Goldie, J. H., & Coldman, A. J. (1998). Drug resistance in cancer: mechanisms and models. Cambridge: Cambridge University Press.

    Book  MATH  Google Scholar 

  • Goldie, J. H., Coldman, A. J., & Gudauskas, G. A. (1982). Rationale for the use of alternating non-cross resistant chemotherapy. Cancer Treat. Rep., 66(3), 439–449.

    Google Scholar 

  • Grantab, R., Sivananthan, S., & Tannock, I. F. (2006). The penetration of anticancer drugs through tumor tissue as a function of cellular adhesion and packing density of tumor cells. Cancer Res., 66(2), 1033–1039.

    Article  Google Scholar 

  • Hakanson, M., Kobel, S., Lutolf, M. P., Textor, M., Cukierman, E., & Charnley, M. (2012). Controlled breast cancer microarrays for the deconvolution of cellular multilayering and density effects upon drug responses. PLoS ONE, 7(6), e40141.

    Article  Google Scholar 

  • Harnevo, L. E., & Agur, Z. (1991). The dynamics of gene amplification described as a multitype compartmental model and as a branching process. Math. Biosci., 103(1), 115–138.

    Article  MathSciNet  MATH  Google Scholar 

  • Jackson, T. L., & Byrne, H. (2000). A mathematical model to study the effects of drug resistance and vasculature on the response of solid tumors to chemotherapy. Math. Biosci., 164(1), 17–38.

    Article  MathSciNet  MATH  Google Scholar 

  • Khain, E., & Sander, L. M. (2006). Dynamics and pattern formation in invasive tumor growth. Phys. Rev. Lett., 96(18), 188103.

    Article  Google Scholar 

  • Kimmel, M., & Axelrod, D. E. (1990). Mathematical models of gene amplification with applications to cellular drug resistance tumorigenicity. Genetics, 125(3), 633–644.

    Google Scholar 

  • Komarova, N. (2006). Stochastic modeling of drug resistance in cancer. J. Theor. Biol., 239(3), 351–366.

    Article  MathSciNet  Google Scholar 

  • Komarova, N., & Wodarz, D. (2005). Drug resistance in cancer: principles of emergence and prevention. Proc. Natl. Acad. Sci. USA, 102(27), 9714–9719.

    Article  Google Scholar 

  • Kreso, A., O’Brien, C. A., van Galen, P., Gal, O. I., Notta, F., et al. (2013). Variable clonal repopulation dynamics influence chemotherapy response in colorectal cancer. Science, 339(6119), 543–548.

    Article  Google Scholar 

  • Laird, A. K. (1964). Dynamics of tumor growth. Br. J. Cancer, 18(3), 490–502.

    Article  Google Scholar 

  • Lavi, O., Gottesman, M. M., & Levy, D. (2012). The dynamics of drug resistance: a mathematical perspective. Drug Resist. Updat., 15(1–2), 90–97.

    Article  Google Scholar 

  • Lavi, O., Greene, J. M., Levy, D., & Gottesman, M. M. (2013). The role of cell density and intratumoral heterogeneity in multidrug resistance. Cancer Res., 73(24), 7168–7175.

    Article  Google Scholar 

  • Lavi, O., Greene, J. M., Levy, D., & Gottesman, M. M. Simplifying the complexity of resistance heterogeneity in metastasis, Trends in Molecular Medicine (2014, accepted).

  • Ledzewicz, U., & Schattler, H. (2006). Drug resistance in cancer chemotherapy as an optimal control problem. Discrete Contin. Dyn. Syst., 6(1), 129–150.

    MathSciNet  MATH  Google Scholar 

  • Long, H., Han, H., Yang, B., & Wang, Z. (2003). Opposite cell density-dependence between spontaneous and oxidative stress-induced apoptosis in mouse fibroblast L-cells. Cell. Physiol. Biochem., 13(6), 401–414.

    Article  Google Scholar 

  • Lorz, A., Mirrahimi, S., & Perthame, B. (2011). Dirac mass dynamics in multidimensional nonlocal parabolic equations. Commun. Partial. Differ. Equ., 36(6), 1071–1098.

    Article  MathSciNet  MATH  Google Scholar 

  • Lorz, A., Lorenzi, T., Hochberg, M., Clairambault, J., & Perthame, B. (2013). Populational adaptive evolution, chemotherapeutic resistance and multiple anti-cancer therapies. Math. Model. Num. Anal., 47(2), 377–399.

    Article  MathSciNet  MATH  Google Scholar 

  • Magal, P., & Webb, G. F. (2000). Mutation, selection and recombination in a model of phenotype evolution. Discrete Contin. Dyn. Syst., 6(1), 221–236.

    MathSciNet  MATH  Google Scholar 

  • Marusyk, A., & Cancer, K. P. (2013). Cancer cell phenotypes, in fifty shades of grey. Science, 339(6119), 528–529.

    Article  Google Scholar 

  • Michelson, S., & Slate, D. (1992). A mathematical model of the P-glycoprotein pump as a mediator of multidrug resistance. Bull. Math. Biol., 54(6), 1023–1038.

    Article  MATH  Google Scholar 

  • Nagane, M., Coufal, F., Lin, H., Bögler, O., Cavenee, W. K., & Huang, H. J. (1996). A common mutant epidermal growth factor receptor confers enhanced tumorigenicity on human glioblastoma cells by increasing proliferation and reducing apoptosis. Cancer Res., 56(21), 5079–5086.

    Google Scholar 

  • Norton, L., & Simon, R. (1977a). Tumor size, sensitivity to therapy, and design of treatment schedules. Cancer Treat. Rep., 61(7), 1307–1317.

    Google Scholar 

  • Norton, L., & Simon, R. (1977b). The growth curve of an experimental tumor following radiotherapy. J. Natl. Cancer Inst., 58(6), 1735–1741.

    Google Scholar 

  • Norton, L., & Simon, R. (1986). The Norton–Simon hypothesis revisited. Cancer Treat. Rep., 70(1), 163–169.

    Google Scholar 

  • Panetta, J. C. (1998). A mathematical model of drug resistance: heterogeneous tumors. Math. Biosci., 147(1), 41–61.

    Article  MATH  Google Scholar 

  • Perthame, B. (2007). Transport equations in biology, frontiers in mathematics. Basel: Birkhäuser.

    MATH  Google Scholar 

  • Perthame, B., & Barles, G. (2008). Dirac concentrations in Lotka–Volterra parabolic PDEs. Indiana Univ. Math. J., 57(7), 3275–3301.

    Article  MathSciNet  MATH  Google Scholar 

  • Qiao, L., & Farrell, G. C. (1999). The effects of cell density, attachment substratum and dexamethasone on spontaneous apoptosis of rate hepatocytes in primary culture. In Vitro Cell. Dev. Biol., Anim., 35(7), 417–424.

    Article  Google Scholar 

  • Saeki, K., Yuo, A., Kato, M., Miyazono, K., Yazaki, Y., & Takaku, F. (1997). Cell density-dependent apoptosis in HL-60 cells, which is mediated by an unknown soluble factor, is inhibited by transforming growth factor beta1 and overexpression of Bcl-2. J. Biol. Chem., 272(32), 20003–20010.

    Article  Google Scholar 

  • Saunders, N. A., Simpson, F., Thompson, E. W., Hill, M. M., Endo-Munoz, L., et al. (2012). Role of intratumoral heterogeneity in cancer drug resistance: molecular and clinical perspectives. EMBO J., 4(8), 675–684.

    Google Scholar 

  • Schuster, R., & Schuster, H. (1995). Reconstruction models for the Ehrlich ascites tumor of the mouse. Math. Pop. Dyn., 2, 335–348.

    Google Scholar 

  • Stein, A. M., Demuth, T., Mobley, D., Berens, M., & Sander, L. M. (2007). A mathematical model of glioblastoma tumor spheroid invasion in a three-dimensional in vitro experiment. Biophys. J., 921, 356–365.

    Article  Google Scholar 

  • Tomasetti, C., & Levy, D. (2010). An elementary approach to modeling drug resistance in cancer. Math. Biosci. Eng., 7(4), 905–918.

    Article  MathSciNet  MATH  Google Scholar 

  • Weaver, V. M., Lelièvre, S., Lakins, J. N., Chrenek, M. A., Jones, J. C., Giancotti, F., Werb, Z., & Bissell, M. J. (2002). Beta4 integrin-dependent formation of polarized three-dimensional architecture confers resistance to apoptosis in normal and malignant mammary epithelium. Cancer Cell, 2(3), 205–216.

    Article  Google Scholar 

  • Zahir, N., & Weaver, V. M. (2004). Death in the third dimension: apoptosis regulation and tissue architecture. Curr. Opin. Genet. Dev., 14(1), 71–80.

    Article  Google Scholar 

Download references

Acknowledgements

We would like to thank George Leiman for editorial assistance. The work was supported by the Intramural Research Program of the National Institutes of Health, Center for Cancer Research, National Cancer Institute and was supported in part by a seed grant from the UMD-NCI Partnership for Cancer Technology. The work of D.L. was supported in part by the joint NSF/NIGMS program under Grant Number DMS-0758374 and by Grant Number R01CA130817 from the National Cancer Institute.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Doron Levy.

Appendix A

Appendix A

Proposition 1

Consider the integro-differential equation (6). If n(x,0)≥0 for all x∈[0,1], then

$$n(x,t) \geq 0 \quad \forall t \in \mathbb{R}_{+}. $$

Proof

The global existence of continuously differentiable solutions of (1) can be obtained following standard arguments (see Perthame, 2007). For the positivity of n, since n≥0 at t=0, due to its continuity, if there exists a t for which n(x,t )=0 for some x∈[0,1] (and n(x,t)>0 for t<t ), then n(y,t )≥0 for all y. Hence, by (6),

$$\begin{aligned} \frac{\partial n(x,t_{*})}{\partial t} = \theta \int_0^1 \! r(y)M(y,x)n(y,t_{*}) \, \mathrm{d} y \geq 0, \end{aligned}$$
(32)

which implies that n(x,t ) is nondecreasing at t . Consequently, n(x,t) cannot pass through 0, as stated. □

Proof of Theorem 1

We follow the proof of Lorz et al. (2013), Lemma 2.2. Consider system (8). If r(x)−c(x)−d(x)<0 for all x∈[0,1], then due to the positivity of n(x,t), \(\frac{\partial n(x,t)}{\partial t} < 0\) for all x∈[0,1]. Hence, \(n(x,t) \xrightarrow[t \rightarrow \infty]{} 0\) in [0,1]. By Lebesgue’s Dominated Convergence Theorem, this implies that ρ(t)→0 as t→∞.

If, on the other hand, there exists x such that r(x )−c(x )−d(x )=0, then these are fixed points of (8), and hence,

$$n(x_{*},t)=n(x_{*},0) \quad \forall t \in \mathbb{R}_{+}. $$

Now suppose that there exists x∈[0,1] such that r(x)−c(x)−d(x)>0. By the continuity of the growth parameters and the compactness of [0,1], r(x)−c(x)−d(x) achieves its maximum, say at \(\{x_{i} \}_{i=1}^{m}\). We note that it is possible that m=∞, or even that the set \(\{x_{i} \}_{i=1}^{m}\) is uncountable (in which case our notation should be altered).

To see that \(\rho(t) \xrightarrow[t \rightarrow \infty]{} \infty\), fix \(x_{j} \in \{x_{i} \}_{i=1}^{m}\) such that n(x j ,0)>0 (as the points where n=0 do not contribute to the growth). Then for all 0<λ<r(x j )−c(x j )−d(x j ), there exists γ λ such that

$$r(x)-c(x)-d(x) \geq \lambda >0 \quad \forall x \in [x_{j}- \gamma_{\lambda},x_{j}+\gamma_{\lambda}]. $$

Let \(h_{\lambda}(t) := \int_{x_{j}-\gamma_{\lambda}}^{x_{j}+\gamma_{\lambda}} \! n(x,t) \, \mathrm{d} x\). Then

$$\frac{d}{dt} h_{\lambda}(t) = \int_{x_{j}-\gamma_{\lambda}}^{x_{j}+\gamma_{\lambda}} \! \bigl[r(x)-c(x)-d(x)\bigr]n(x,t) \, \mathrm{d} x \geq \lambda \int _{x_{j}-\gamma_{\lambda}}^{x_{j}+\gamma_{\lambda}} \! n(x,t) \, \mathrm{d} x = \lambda h_{\lambda} (t), $$

so that, for a positive constant h(0),

$$\begin{aligned} h_{\lambda}(t) \geq h(0)e^{\lambda t}. \end{aligned}$$
(33)

As \(\rho(t) = \int_{0}^{1} \! n(x,t) \, \mathrm{d} x \geq h_{\lambda}(t)\), (33) implies that \(\rho(t) \xrightarrow[t \rightarrow \infty]{} \infty\), as desired. To find the limiting distribution, note that for \(x \notin \{x_{i} \}_{i=1}^{m}\), choose λ such that

$$r(x)-c(x)-d(x) < \lambda < r(x_{j})-c(x_{j})-d(x_{j}). $$

Then,

$$\begin{aligned} \frac{n(x,t)}{\rho(t)}=\frac{n(x,0)e^{[r(x)-c(x)-d(x)]t}}{\rho(t)} \leq \frac{n(x,0)}{h(0)} e^{[r(x)-c(x)-d(x)-\lambda]t} \xrightarrow[t \rightarrow \infty]{} 0. \end{aligned}$$
(34)

As \(\int_{0}^{1} \! \frac{n(x,t)}{\rho(t)} \, \mathrm{d} x = 1\), for all time t, we have the desired result, namely

$$\lim_{t \rightarrow \infty} \frac{n(x,t)}{\rho(t)}=\sum _{i=1}^{m} a_{i} \delta(x-x_{i}), $$

with \(\sum_{i=1}^{m} a_{i} = 1\). If the number of maximizers is uncountable, a similar result will hold for a continuous measure. □

Proof of Theorem 2

Let n(x,t) satisfy (16). By Proposition 1, since n(x,t)≥0, ρ(t)≥0 as well. Let r M ,c m , and d m be constants such that r(x)≤r M , c(x)≥c m >0, and d(x)≥d m >0, and recall that G≥0. Hence, ρ′(t) can be bounded above by

$$\frac{d \rho(t)}{dt} = \int_0^1 \! \bigl[r(x)-c(x)-G\bigl(\rho(t)\bigr)d(x)\bigr]n(x,t) \, \mathrm{d} x \leq \bigl[r_{M}-c_{m}-G\bigl(\rho(t)\bigr)d_{m}\bigr] \rho(t). $$

Since \(G(\rho) \xrightarrow[\rho \rightarrow \infty]{} \infty\), there exists ρ M such that ρ(0)≤ρ M and r M c m G(ρ M )d m <0. This implies that at ρ=ρ M , ρ′(t)<0, and hence ρ(t)≤ρ M . □

Rights and permissions

Reprints and permissions

About this article

Cite this article

Greene, J., Lavi, O., Gottesman, M.M. et al. The Impact of Cell Density and Mutations in a Model of Multidrug Resistance in Solid Tumors. Bull Math Biol 76, 627–653 (2014). https://doi.org/10.1007/s11538-014-9936-8

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11538-014-9936-8

Keywords

Navigation