Skip to main content
Log in

Propagation of Chaos in the Nonlocal Adhesion Models for Two Cancer Cell Phenotypes

  • Published:
Journal of Nonlinear Science Aims and scope Submit manuscript

Abstract

We establish a quantitative propagation of chaos for a large stochastic systems of interacting particles. We rigorously derive a mean-field system, which is a diffusive cell-to-cell nonlocal adhesion model for two different phenotypes of tumors, from that stochastic system as the number of particles tends to infinity. We estimate the error between the solutions to a N-particle Liouville equation associated with the particle system and the limiting mean-field system by employing the relative entropy argument.

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.

Similar content being viewed by others

Notes

  1. In Ahn et al. (2021), the adhesion velocities for u and v are given by \( \int _{E(x)}\left[ M_{11}u(x+y,t)+ M_{12}v(x+y,t) \right] \omega (y)\,dy \) and \( \int _{E(x)}\left[ M_{21}u(x+y,t)+ M_{22}v(x+y,t) \right] \omega (y)\,dy, \) respectively. Here E(x) is a sensing domain depending on x in a bounded domain \({ \Omega }\) and \(\omega \) is a given adhesion strength function.

  2. Let the \((T_n)_{n\ge 0}\) be a strictly increasing sequence of positive random variables with \(T_0 =0\). The nonnegative integer valued process N defined by \(N_t = \sum _{n\ge 1} 1_{\{ t \ge T_n\}}\) is called the counting process associated to the sequence \((T_n)_{n\ge 1}\).

References

  • Ahn, J., Chae, M., Lee, J.: Nonlocal adhesion models for two cancer cell phenotypes in a multidimensional bounded domain. Z. Angew. Math. Phys. 72, 48 (2021)

    Article  MathSciNet  MATH  Google Scholar 

  • Albi, G., Bongini, M., Rossi, F., Solombrino, F.: Leader formation with mean-field birth and death models. Math. Models Methods Appl. Sci. 29, 633–679 (2019)

    Article  MathSciNet  MATH  Google Scholar 

  • Anderson, A.R.A., Chaplain, M.A.J., Newman, E.L., Steele, R.J.C., Thompson, A.M.: Mathematical modelling of tumour invasion and metastasis. J. Theor. Med. 2, 129–154 (2000)

    Article  MATH  Google Scholar 

  • Anibal, R.-B.: The heat equation with general periodic boundary conditions. Potential Anal. 46, 295–321 (2017)

    Article  MathSciNet  MATH  Google Scholar 

  • Armstrong, N.J., Painter, K.J., Sherratt, J.A.: A continuum approach to modelling cell-cell adhesion. J. Theor. Biol. 243, 98–113 (2006)

    Article  MathSciNet  MATH  Google Scholar 

  • Baudoin, F.: Diffusion Processes and Stochastic Calculus, EMS, (2014)

  • Bellomo, N., Li, N.K., Maini, P.K.: On the foundations of cancer modelling: selected topics, speculations, and perspectives. Math. Models Methods Appl. Sci. 18, 593–646 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  • Bitsouni, V., Chaplain, M.A.J., Eftimie, R.: Mathematical modelling of cancer invasion: the multiple roles of TGF-\(\alpha _2\) pathway on tumour proliferation and cell adhesion. Math. Models Methods Appl. Sci. 27, 1929–1962 (2017)

    Article  MathSciNet  MATH  Google Scholar 

  • Bolley, F., Canizo, J.A., Carrillo, J.A.: Stochastic mean-field limit: non-Lipschitz forces and swarming. Math. Models Methods Appl. Sci. 21, 2179–2210 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  • Brenier, Y.: Convergence of the Vlasov-Poisson system to the incompressible Euler equations. Commun. Partial Differ. Equ. 25, 737–754 (2000)

    Article  MathSciNet  MATH  Google Scholar 

  • Bresch, D., Jabin, P.-E., Wang, Z.: Mean-field limit and quantitative estimates with singular attractive kernels, preprint

  • Carrillo, J.A., Choi, Y.-P.: Mean-field limits: from particle descriptions to macroscopic equations. Arch. Ration. Mech. Anal. 241, 1529–1573 (2021)

    Article  MathSciNet  MATH  Google Scholar 

  • Carrillo, J.A., Choi, Y.-P., Jung, J.: Quantifying the hydrodynamic limit of Vlasov-type equations with alignment and nonlocal forces. Math. Models Methods Appl. Sci. 31, 327–408 (2021)

    Article  MathSciNet  MATH  Google Scholar 

  • Carrillo, J.A., Choi, Y.-P., Peng, Y.: Large friction-high force fields limit for the nonlinear Vlasov-Poisson-Fokker-Planck system. Kinet. Relat. Models 15, 355–384 (2022)

    Article  MathSciNet  Google Scholar 

  • Carrillo, J.A., Choi, Y.-P., Salem, S.: Propagation of chaos for the Vlasov-Poisson-Fokker-Planck equation with a polynomial cut-off. Commun. Contemp. Math. 21, 1850039 (2019)

    Article  MathSciNet  MATH  Google Scholar 

  • Chaplain, M.A.J., Anderson, A.R.A.: Mathematical modelling of tissue invasion. In: Preziosi, L. (ed.) Cancer Modelling and Simulation, pp. 267–297. Chapman Hall, London (2003)

    Google Scholar 

  • Chaplain, M.A.J., Lolas, G.: Mathematical modelling of cancer invasion of tissue: the role of the urokinase plasminogen activation system. Math. Models Methods Appl. Sci. 15, 1685–1734 (2005)

    Article  MathSciNet  MATH  Google Scholar 

  • Choi, Y.-P.: Large friction limit of pressureless Euler equations with nonlocal forces. J. Differ. Equ. 299, 196–228 (2021)

    Article  MathSciNet  MATH  Google Scholar 

  • Choi, Y.-P., Jeong, I.-J.: Relaxation to the fractional porous medium equation from the Euler-Riesz system. J. Nonlinear Sci. 2, 31 (2021)

    MathSciNet  MATH  Google Scholar 

  • Choi, Y.-P., Jung, J.: Modulated energy estimates for singular kernels and its applications to asymptotic analyses for kinetic equations, preprint

  • Choi, Y.-P., Kang, K., Kim, H.K., Kim, J.-M.: Temporal decays and asymptotic behaviors for a Vlasov equation with a flocking term coupled to incompressible fluid flow. Nonlinear Anal. Real World Appl. 63, 103410 (2022)

    Article  MathSciNet  MATH  Google Scholar 

  • Choi, Y.-P., Salem, S.: Propagation of chaos for aggregation equations with no-flux boundary conditions and sharp sensing zones. Math. Models Methods Appl. Sci. 28, 223–258 (2018)

    Article  MathSciNet  MATH  Google Scholar 

  • Choi, Y.-P., Salem, S.: Collective behavior models with vision geometrical constraints: truncated noises and propagation of chaos. J. Differ. Equ. 266, 6109–6148 (2019)

    Article  MathSciNet  MATH  Google Scholar 

  • Dafermos, C.M.: The second law of thermodynamics and stability. Arch. Ration. Mech. Anal. 70, 167–179 (1979)

    Article  MathSciNet  MATH  Google Scholar 

  • Dellacherie, C., Meyer, P.-A.: Probabilities and potential. North-Holland Mathematics Studies, vol. 29. North-Halland Publishing Co., Amsterdam (1978)

  • Duerinckx, M.: Mean-field limits for some Riesz interaction gradient flows. SIAM J. Math. Anal. 48, 2269–2300 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  • Dupuis, P., Ellis, R.S.: A Weak Convergence Approach to the Theory of Large Deviations. Wiley Series in Probability and Statistics: Probability and Statistics, Wiley, New York (1997)

    Book  MATH  Google Scholar 

  • Durrett, R.: Propability Theory and Examples. Cambridge University Press, Cambridge (2010)

    Book  MATH  Google Scholar 

  • Hauray, M., Salem, S.: Propagation of chaos for the Vlasov-Poisson-Fokker-Planck system in 1D. Kinet. Relat. Models 12, 269–302 (2019)

    Article  MathSciNet  MATH  Google Scholar 

  • Huang, H., Liu, J.-G., Pickl, P.: On the mean-field limit for the Vlasov-Poisson-Fokker-Planck system. J. Stat. Phys. 181, 1915–1965 (2020)

    Article  MathSciNet  MATH  Google Scholar 

  • Jabin, P.-E., Wang, Z.: Mean field limit and propagation of chaos for Vlasov systems with bounded forces. J. Funct. Anal. 271, 3588–3627 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  • Jabin, P.-E., Wang, Z.: Mean field limit for stochastic particle systems, Active particles. Vol. 1. Advances in theory, models, and applications, pp. 379–402, Birkhäuser/Springer, Cham, (2017)

  • Jabin, P.-E., Wang, Z.: Quantitative estimates of propagation of chaos for stochastic systems with \(W^{-1,\infty }\) kernels. Invent. Math. 214, 523–591 (2018)

    Article  MathSciNet  MATH  Google Scholar 

  • Jeanblanc, M.: Jump Processes, lecture note distributed in CIMPA School, Marrakesh (2007)

  • Kac, M.: Foundations of kinetic theory, in Proceedings of the Third Berkeley Symposium on Mathematical Statistics and Probability, 1954–1955, vol. III, University of California Press, Berkeley, pp. 171–197 (1956)

  • Klebaner, F.C.: Introduction to Stochastic Calculus with Applications, 2nd edn. Imperial College Press, London (2005)

    Book  MATH  Google Scholar 

  • Kolbe, N., Lukáčová-Medvid’ová, M., Sfakianakis, N., Wiebe, B.: Numerical simulation of a contractivity based multiscale cancer invasion model. In: Gerisch, A., Penta, R., Lang, J. (eds.) Multiscale Models in Mechano and Tumor Biology. Series Title: Lecture Notes in Computational Science and Engineering, vol. 122, pp. 73–91. Springer, Cham (2017)

    Chapter  Google Scholar 

  • Kolbe, N., Sfakianakis, N., Stinner, C., Surulescu, C., Lenz, J.: Modeling multiple taxis: tumor invasion with phenotypic heterogeneity, haptotaxis, and unilateral interspecies repellence. Discrete Cont. Dyn. B. 26, 443–481 (2021)

    MathSciNet  MATH  Google Scholar 

  • Kurtz, T.G.: Limit theorems for sequences of jump Markov processes approximating ordinary differential processes. J. Appl. Probability 8, 344–356 (1971)

    Article  MathSciNet  MATH  Google Scholar 

  • Lattanzio, C., Tzavaras, A.E.: From gas dynamics with large friction to gradient flows describing diffusion theories. Commun. Partial Differ. Equ. 42, 261–290 (2017)

    Article  MathSciNet  MATH  Google Scholar 

  • Lim, T.S., Lu, Y., Nolen, J.H.: Quantitative propagation of chaos in a bimolecular chemical reaction-diffusion model. SIAM J. Math. Anal. 52, 2098–2133 (2020)

    Article  MathSciNet  MATH  Google Scholar 

  • Litcanu, G., Morales-Rodrigo, C.: Asymptotic behavior of global solutions to a model of cell invasion. Math. Models Methods Appl. Sci. 20, 1721–1758 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  • Morales-Rodrigo, C.: Local existence and uniqueness of regular solutions in a model of tissue invasion by solid tumours. Math. Comput. Modell. 47, 604–613 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  • Nguyen, Q.H., Rosenzweig, M., Serfaty, S.: Mean-field limits of Riesz-type singular flows with possible multiplicative transport noise, preprint

  • Oelschlager, K.: On the derivation of reaction-diffusion equations as limit dynamics of systems of moderately interacting stochastic processes. Probab. Theory Related Fields 82, 565–586 (1989)

    Article  MathSciNet  MATH  Google Scholar 

  • Perumpanani, A.J., Byrne, H.M.: Extracellular matrix concentration exerts selection pressure on invasive cells. Eur. J. Cancer 35, 1274–1280 (1999)

    Article  Google Scholar 

  • Protter, P.E.: Stochastic Intergration and Differential Equations, 2nd edn. Springer, Berlin (2004)

    Google Scholar 

  • Saint-Raymond, L.: Hydrodynamic Limits of the Boltzmann Equation. Lecture Notes in Mathematics, vol. 1971. Springer, Berlin (2009)

  • Salem, S.: Propagation of chaos for fractional Keller Segel equations in diffusion dominated and fair competition cases. J. Math. Pures Appl. 132, 79–132 (2019)

    Article  MathSciNet  MATH  Google Scholar 

  • Schönbucher, P.J.: Credit Derivatives Pricing Models. Wiley, New York (2003)

    Google Scholar 

  • Serfaty, S.: Mean field limits of the Gross-Pitaevskii and parabolic Ginzburg-Landau equations. J. Am. Math. Soc. 30, 713–768 (2017)

    Article  MathSciNet  MATH  Google Scholar 

  • Serfaty, S.: Mean field limit for Coulomb-type flows. Duke Math. J. 169, 2887–2935 (2020)

    Article  MathSciNet  MATH  Google Scholar 

  • Stinner, C., Surulescu, C., Winkler, M.: Global weak solutions in a PDE-ODE system modeling multiscale cancer cell invasion. SIAM J. Math. Anal. 46, 1969–2007 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  • Sznitman, A.-S.: Topics in propagation of chaos. In: École d’Éte de Probabilités de Saint-Flour XIX–1989, Lecture Notes in Mathematics, Vol. 1464, pp. 165–251 (Springer, 1991)

  • Tao, Y., Wang, M.: Global solution for a chemotactic-haptotactic model of cancer invasion. Nonlinearity 21, 2221–2238 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  • Tao, Y., Winkler, M.: A critical virus production rate for blow-up suppression in a haptotaxis model for oncolytic virotherapy. Nonlinear Anal. 198, 111870 (2020)

    Article  MathSciNet  MATH  Google Scholar 

  • Tao, Y., Winkler, M.: Global classical solutions to a doubly haptotactic cross-diffusion system modeling oncolytic virotherapy. J. Differ. Equ. 268, 4973–4997 (2020)

    Article  MathSciNet  MATH  Google Scholar 

  • Yau, H.-T.: Relative entropy and hydrodynamics of Ginzburg-Landau models. Lett. Math. Phys. 22, 63–80 (1991)

    Article  MathSciNet  MATH  Google Scholar 

  • Yeung, K.T., Yang, J.: Epithelial-mesenchymal transition in tumor metastasis. Mol. Oncol. 11, 28–39 (2017)

    Article  Google Scholar 

Download references

Acknowledgements

We would like to sincerely thank the anonymous referee for helpful comments and suggestions. We thank Guang Yang for helpful conversations about showing the non-positivity of (B.2) based on probabilistic arguments. J. Ahn was supported by the Dongguk University Research Fund of 2020. M. Chae was supported by NRF-2018R1A1A3A04079376. Y.-P. Choi was supported by National Research Foundation of Korea(NRF) grant funded by the Korea government(MSIP) (No. 2017R1C1B2012918 and 2022R1A2C1002820) and Yonsei University Research Fund of 2021-22-0301. J. Lee is supported by SSTF-BA1701-05 (Samsung Science and Technology Foundation).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Myeongju Chae.

Additional information

Communicated by Philipp M Altrock.

Publisher's Note

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

The manuscript I submitted has no associated data.

Appendices

Appendix A. Poisson Jump Process

In this appendix, we briefly introduce the Poisson jump process and its properties used in the paper. We refer to Durrett (2010), Jeanblanc (2007), Protter (2004) and (Schönbucher 2003, Chapter 4). For the basic definitions for stochastic processes we refer to (Baudoin 2014; Protter 2004).

Let \((\Omega , { {\mathcal {F}} }, P)\) be a complete probability space with a filtration \(({ {\mathcal {F}} }_t)_{t\ge 0}\).

Definition 2

A nonnegative integer valued stochastic process \(E=(E_t)_{t\ge 0}\) defined on \((\Omega , { {\mathcal {F}} }, P)\) is called Poisson jump process with intensity or rate \(\lambda \) if the process has independent and stationary increments:

  1. (i)

    \(E_0=0\) a.s.

  2. (ii)

    \(E_{t_k} - E_{t_{k-1}}\) \((k=1, \dots , n)\) are independent for \(0= t_0< t_1<\cdots < t_n\).

  3. (iii)

    \(E_t-E_s\) is \(Poisson(\lambda (t-s))\), that is,

    $$\begin{aligned} P(E_t-E_s = k)= e^{-\lambda (t-s)} \lambda ^k (t-s)^k/k! \, \text{ for } k=0, 1, 2, \ldots . \end{aligned}$$

\(E_t\) is thought as a number of events occurring in [0, t] with the rate \(\lambda \). It satisfies \( {\mathbb {E}} E_t = \lambda t\), \(\mathrm{Var} E_t = \lambda t\). From the independence increments property we find \(E_t-\lambda t\) is a martingale such that

$$\begin{aligned} {\mathbb {E}} (E_t-E_s| { {\mathcal {F}} }) = \lambda (t-s). \end{aligned}$$

The Poisson process with constant intensity \(\lambda \) is generalized to a counting processes \(N=(N_t)_{t\ge 0}\)Footnote 2 with a predictable compensator process \(A=(A_t)_{t\ge 0}\), for which the compensated process \(N_t-A_t\) is a local martingale. The compensator can give an information about the probabilities of jumps over the next time step:

$$\begin{aligned} {\mathbb {E}} (N_{t+ h} -N_t | { {\mathcal {F}} }_t) = {\mathbb {E}}(A_{t+ h}- A_t| { {\mathcal {F}} }_t) \end{aligned}$$

if we restrict to the case the compensated process is a true martingale. If A is differentiable, we extend the definition of (homogeneous) intensity \(\lambda \) for \(E_t\) to the intensity process \(\lambda _t\) for \(N_t\):

$$\begin{aligned} A_t= \int _0^t \lambda _s\, ds. \end{aligned}$$

Definition 3

Let \(\lambda \) be a nonnegative \(({ {\mathcal {F}} })_{t\ge 0}\) adapted process defined on \(({ \Omega }, { {\mathcal {F}} }, P)\) such that \(\int _0^t \lambda _s\,ds <\infty .\) A counting process N is said to be an inhomogeneous Poisson process with stochastic intensity \(\lambda \) if the process

$$\begin{aligned}M_t = N_t - \int _0^t \lambda _s\,ds, \quad t\ge 0\end{aligned}$$

is a martingale.

In Sect. 2.1, we used that \(\alpha _k \chi _1(\Xi ^i_{s^-})\) (\(k=1,2\)) are intensities of the time-changed Poisson processes \({\tilde{E}}_k^i(t) = E^i(\alpha _k \int _0^t \chi _k(\Xi ^i_{s^-}) \,ds\) at time t (Oelschlager 1989) to see the expectation of the stochastic integral with respect to \({\tilde{E}}_k^i(t)\) is as same as that with respect to \(A_k^i(t):=\alpha _k \int _0^t \chi _1(\Xi ^i_{s^-})\,ds\) (\(k=1,2\)).

Appendix B. Proof of (3.5)

In this appendix, we prove (3.5) following the similar argument in (Lim et al. 2020, Lemma 4.1). Although the overall idea is similar, we add more details in this probabilistic argument for reader’s convenience.

We note that the process \((\mathbf {X}_{t}^{N}, \mathbf {\Xi }_{t}^{N})\) given by (1.2)–(1.3) is a Markov process since it is a solution to a SDE driven by Markov processes \(\{B_t^i\}_{t\ge 0}, \{E_t^i\}_{t\ge 0}\) for \(i=1, \dots , N\). We refer to (Protter 2004, Theorem 32, Chapter 5) for the Markov nature of the process defined by stochastic integral against Lévy processes. Fix \(t_0>0\) and we will show that (3.5) holds at \(t=t_0\). Let u and \({\tilde{u}}\) be the solutions of the following equations:

$$\begin{aligned} \begin{aligned} \partial _t u&= { {\mathcal {L}} }^*_{r,N} u \quad \text{ with } u(t_0) = \rho _N(t_0), \\ \partial _t {\tilde{u}}&= { {\mathcal {L}} }^*_{r,N} {\tilde{u}} \quad \text{ with } {\tilde{u}}(t_0) = {\bar{\rho }}_N(t_0). \end{aligned} \end{aligned}$$
(B.1)

By Theorem 2 in Sect. 4, u and \({\tilde{u}}\) are in \(C([t_0, \infty ); W^{1, p}(\Pi ^N))\) (\(F=0\) case) with initial data \(\rho _N(t_0), {\bar{\rho }}_N(t_0) \in W^{1, p}(\Pi ^N)\). We find that

$$\begin{aligned} \frac{d}{dt}{ {\mathcal {H}} }(u|{\tilde{u}})(t) = \int _{\Pi ^N} \left( \log \left( \frac{u(t)}{{\tilde{u}}(t) }\right) { {\mathcal {L}} }^*_{r,N} u(t)- u(t) \left( \frac{{ {\mathcal {L}} }^*_{r,N} u(t)}{u(t)} - \frac{{ {\mathcal {L}} }^*_{r,N} \tilde{u}(t)}{{\tilde{u}}(t)}\right) \right) dm^N\nonumber \\ \end{aligned}$$
(B.2)

for \(t> t_0\). Let us assume the non-positivity of (B.2) for the moment. Note that the LHS of (3.5) is formally obtained if we evaluate \(\frac{d}{dt}{ {\mathcal {H}} }(u|{\tilde{u}})(t)\) at \(t=t_0\). Indeed, we see that (B.2) converges to

$$\begin{aligned} \int _{\Pi ^N} \left( \log \left( \frac{\rho _N}{{\bar{\rho }}_N} \right) { {\mathcal {L}} }^*_{r,N} \rho _N - \rho _N \left( \frac{{ {\mathcal {L}} }^*_{r,N} \rho _N}{\rho _N} - \frac{{ {\mathcal {L}} }^*_{r,N} {\bar{\rho }}_N}{{\bar{\rho }}_N}\right) \right) dm^N \end{aligned}$$
(B.3)

as \(t \searrow t_0\); the difference of two integrals (B.2) and (B.3) is

$$\begin{aligned}&\int _{\Pi ^N} \left( ({ {\mathcal {L}} }^*_{r,N} u - { {\mathcal {L}} }^*_{r,N} \rho _N) \log \frac{u}{{\tilde{u}}} + { {\mathcal {L}} }^*_{r,N}\rho _N (\log u-\log \rho _N - (\log {\tilde{u}} - \log {\bar{\rho }}_N))\right) dm^N\nonumber \\&\quad + \int _{\Pi ^N} \frac{u(t)}{{\tilde{u}}(t)} ( { {\mathcal {L}} }^*_{r,N} {\tilde{u}} - { {\mathcal {L}} }^*_{r,N} {\bar{\rho }}_N) + { {\mathcal {L}} }^*_{r,N}\bar{\rho }_N\left( \frac{u}{{\tilde{u}}}- \frac{\rho _N}{\bar{\rho }_N}\right) dm^N. \end{aligned}$$
(B.4)

For the moment we assume the positivity of solutions \(\rho _N\) and \({\bar{\rho }}_N\) to (3.1) and (3.2) respectively, which will be justified in Lemma 3 below. By integrating by parts, we have

$$\begin{aligned}(B.4)&=-\frac{1}{2}\int _{\Pi ^N} (\nabla _N u - \nabla _N \rho _N) \cdot \left( \frac{\nabla _N u }{u} - \frac{\nabla _N {\tilde{u}}}{{\tilde{u}}} \right) dm^N \\&\quad +\int _{\Pi ^N} (S_N^* u-S_N^* \rho _N) (\log u - \log {\tilde{u}})\,dm^N \\&\quad - \frac{1}{2} \int _{\Pi ^N} \nabla _N \rho _N\cdot \left( \frac{\nabla _N u}{u} - \frac{\nabla _N \rho _N}{\rho _N}- \left( \frac{\nabla _N {\tilde{u}}}{{\tilde{u}}} - \frac{\nabla _N {\bar{\rho }}_N}{{\bar{\rho }}_N} \right) \right) dm^N \\&\quad + \int _{\Pi ^N} S_N^* \rho _N ( \log u - \log \rho _N - (\log {\tilde{u}} - \log {\bar{\rho }}_N))\,dm^N \\&\quad -\frac{1}{2} \int _{\Pi ^N} \left( \nabla _N \left( \frac{u}{\tilde{u}}\right) \cdot (\nabla _N {\tilde{u}} - \nabla _N {\bar{\rho }}_N) - 2\frac{u}{{\tilde{u}}} ( S_N^* {\tilde{u}} - S_N^* {\bar{\rho }}_N) \right) dm^N \\&\quad -\frac{1}{2} \int _{\Pi ^N} \left( \nabla _N {\bar{\rho }}_N \cdot \nabla _N \left( \frac{ u}{ {\tilde{u}} }- \frac{\rho _N}{ {\bar{\rho }}_N}\right) - 2 S_N^* {\bar{\rho }}_N \left( \frac{ u}{ {\tilde{u}} }- \frac{\rho _N}{ {\bar{\rho }}_N} \right) \right) dm^N . \end{aligned}$$

Using \( u, {\tilde{u}}, \rho _N, {\bar{\rho }}_N >0\) and \(u, {\tilde{u}} \in C([t_0, \infty ); W^{1, p}(\Pi ^N))\) for \(p>dN\), we see that the difference is bounded above by \(C ( \Vert u(t) - \rho _N\Vert _{W^{1, p}(\Pi ^N)} + \Vert {\tilde{u}}(t) - {\bar{\rho }}_N\Vert _{W^{1, p}(\Pi ^N)} )\), which is vanishing as \( t \searrow t_0\). It concludes (3.5).

Proof of the non-positivity of (B.2). In what follows, we suppress the superscript N in \( \mathbf {Y_t^N}=(\mathbf {X}_{t}^{N}, \mathbf {\Xi }_{t}^{N})\). Let \(\mathbf {Y}_t \), \({\tilde{\mathbf {Y}}}_t \) be the Markov process defined by (1.2)–(1.3) with initial distribution \(\rho _N\), \({\bar{\rho }}_N\) at \(t=t_0\) respectively. We denote the solutions of (B.1) at time t by \(u_t\) and \({\tilde{u}}_t\), which are distribution of \(\mathbf {Y}_t \), \({\tilde{\mathbf {Y}}}_t \). Let \(\kappa \) be the Borel measure on \(\Pi \times \Pi \) induced by the joint distribution of \((\mathbf {Y}_t, \mathbf {Y}_{t+h})\) for \(t>t_0, h>0\). That is

$$\begin{aligned} \kappa (B) = P( (\mathbf {Y}_t, \mathbf {Y}_{t+h}) \in B )\quad \text{ for } \text{ any } B\in { {\mathcal {B}} }(\Pi \times \Pi ), \end{aligned}$$

where \({ {\mathcal {B}} }(\Pi \times \Pi )\) is the collection of Borel sets in \(\Pi \times \Pi \). We denote the induced Borel measure on \(\Pi \) by \(\mathbf {Y}_t\) by \(\lambda \). The induced measures \({\tilde{\kappa }}\), \(\tilde{\lambda }\) are similarly defined.

Since the space \(\Pi \) is sufficiently nice, it is possible to disintegrate \(\kappa \) and \({\tilde{\kappa }}\). More precisely, we can find a kernel \(\lambda \) (\({\tilde{\lambda }}\)) -a.e. uniquely determined kernel \(\mu _{\cdot }\) (\({\tilde{\mu }}_{\cdot }\)) that maps \(\Pi \) to the space of Borel measures on \(\Pi \) such that

  1. (i)

    For \(x \in \Pi \), \(\mu _x( \cdot )\) is a Borel measure on \(\Pi \).

  2. (ii)

    \(\mu _{\cdot }(B)\) is \(\lambda \)-measurable for any fixed \(B \in { {\mathcal {B}} }(\Pi )\).

  3. (iii)

    \(\int _{\Pi \times \Pi } f(x, y) d\kappa (x, y) = \int _{\Pi }\int _{\Pi } f(x,y) \mu _x (dy) \lambda (dx)\) for any bounded Borel function on \(\Pi \times \Pi \).

The similar assertion holds for \({\tilde{\mu }}_{\cdot }\). The disintegration theorem holds in more general circumstance, but we won’t go further, see for instance (Dellacherie and Meyer 1978, III-70). By (i)–(iii), we have \(\{\mu _x (dy)\}_{x\in \Pi }\) and \(\{{\tilde{\mu }}_x (dy)\}_{x\in \Pi }\) such that

$$\begin{aligned} P( \mathbf {Y}_t \in A, \mathbf {Y}_{t+h} \in B)&= \int _A\int _B \mu _x(dy) d\lambda (x) \\ P({\tilde{\mathbf {Y}}}_t \in A, {\tilde{\mathbf {Y}}}_{t+h} \in B)&= \int _A\int _B {\tilde{\mu }}_x(dy) d{\tilde{\lambda }}(x) \end{aligned}$$

for \(A, B \in { {\mathcal {B}} }(\Pi )\). On the other hands, since \(\{ \mathbf {Y}_t\}\), \(\{{\tilde{\mathbf {Y}}}_t\}\) are the same Markov processes, there exists the transition kernel \(P_h(x, dy)\) such that

$$\begin{aligned} P( \mathbf {Y}_t \in A, \mathbf {Y}_{t+h} \in B)&= \int _A\int _B P_h(x, dy) d\lambda (x), \\ P({\tilde{\mathbf {Y}}}_t \in A, {\tilde{\mathbf {Y}}}_{t+h} \in B)&= \int _A\int _B P_h(x, dy) d{\tilde{\lambda }}(x). \end{aligned}$$

Thus we conclude

$$\begin{aligned} \mu _x (dy) = P_h(x, dy) = {\tilde{\mu }}_x (dy). \end{aligned}$$

We use \(u_{t+h| t}(x, y)\) to denote the density of \(\mu _x (dy)\) with respect to dy, where we abuse the notation dy to stand for \(dm=dy \otimes d \sharp \) the canonical measure on \(\Pi = {\mathbb {T}}^d \times \{ 1, 2\}\). \({\tilde{u}}_{t+h| t}(x, y)\) is defined in the same way. \(u_{t+h| t}(x, y)\) is so called a conditional density, which is a density of conditional probability

$$\begin{aligned} P(u_{t+h} \in B|u_t = x) = \int _B \mu _x (dy). \end{aligned}$$

Note that \(u_t\) and \(u_{t+h}\) are the x- and y- marginal densities of the joint distribution of \((Y_t, Y_{t+h})\), while the density of \(\kappa \) is denoted by \(u_{t, t+h}\). The notations with tilde are defined accordingly. As in (Lim et al. 2020, Lemma 4.1), we define the average conditional relative entropy between \(u_{t+h|t}\) and \({\tilde{u}}_{t+h|t}\) by

$$\begin{aligned} { {\mathcal {H}} }( u_{t+h|t} | {\tilde{u}}_{t+h|t}): = \int _{\Pi } u_t(x) \int _{\Pi } u_{t+h|t}(x,y) \log \frac{ u_{t+h|t}}{{\tilde{u}}_{t+h|t}} \,dydx. \end{aligned}$$

Applying the chain rule (Dupuis and Ellis 1997, Theorem C.3.1) of the relative entropy, we have

$$\begin{aligned} { {\mathcal {H}} }( u_{t, t+h} | {\tilde{u}}_{t, t+h}) = { {\mathcal {H}} }( u_{t} | \tilde{u}_{t})+{ {\mathcal {H}} }( u_{t+h|t} | {\tilde{u}}_{t+h|t}). \end{aligned}$$

The last quantity is zero since \(u_{t+h|t} = {\tilde{u}}_{t+h|t}\). If we proceed the above discussion to the joint distribution of \((\mathbf {Y}_{t+h}, \mathbf {Y}_t)\), we similarly arrive at

$$\begin{aligned} { {\mathcal {H}} }( u_{ t+h,t} | {\tilde{u}}_{t+h,t}) = { {\mathcal {H}} }( u_{t+h} | \tilde{u}_{t+h})+{ {\mathcal {H}} }( u_{t|t+h} | {\tilde{u}}_{t|t+h}). \end{aligned}$$

Finally using \({ {\mathcal {H}} }( u_{ t+h,t} | {\tilde{u}}_{t+h,t}) ={ {\mathcal {H}} }( u_{t, t+h} | {\tilde{u}}_{t, t+h})\) and the non-negativity of the relative entropy \({ {\mathcal {H}} }( u_{t|t+h} | {\tilde{u}}_{t|t+h}) \), we have

$$\begin{aligned} { {\mathcal {H}} }( u_{t} | {\tilde{u}}_{t}) \ge { {\mathcal {H}} }( u_{t+h} | \tilde{u}_{t+h}), \end{aligned}$$

thus (B.2) \(\le 0\).

In the lemma below we provide the positivity of \(\rho _N\) and \({\bar{\rho }}_N\) to (3.1) and (3.2) respectively, which was assumed in showing (B.4) to vanish.

Lemma 3

Let \(\rho _N\) and \({\bar{\rho }}_N\) be the solutions of (3.1) and (3.2) respectively, which well-posedness are given in Theorem 2 in Sect. 4 with initial data \(\rho _N(0), {\bar{\rho }}_N(0) \in W^{1, p}(\Pi ^N)\) for \(p>dN\). If the initial data \(\rho _N(0)\) and \({\bar{\rho }}_N(0)\) satisfy

$$\begin{aligned} \inf _{(\mathbf {x},\varvec{\xi }) \in \Pi ^N}\rho _N(0)> 0 \quad \text{ and } \quad \inf _{(\mathbf {x},\varvec{\xi }) \in \Pi ^N}{\bar{\rho }}_N(0) > 0, \end{aligned}$$

then the solutions \(\rho _N\) and \({\bar{\rho }}_N\) remain positive.

Proof

We only show this for \(\rho _N\) since almost the same argument can be applied for \({\bar{\rho }}_N\). Noticing \(\xi _i \in \{1,2\}\) for all \(i=1,\dots ,N\), for the proof, we use the induction on m, based on the number of cases \(\xi _i = 2\). Let us start with \(m=0\), i.e., \(\xi _i = 1\) for all \(i=1,\dots ,N\). In this case, we observe (2.5) to read

$$\begin{aligned} \partial _t \rho _N&= (\Delta _N + F^*_N)\rho _N \\&\quad + \alpha _1 \sum _{i=1}^N \left( \chi _2(\xi _i)\rho _N(\mathbf {x}, \tilde{\Theta }_2^i(\varvec{\xi }))- \chi _1(\xi _i)\rho _N(\mathbf {x}, \varvec{\xi }) \right) \\&\quad +\alpha _2 \sum _{i=1}^N\left( \chi _1(\xi _i)\rho _N(\mathbf {x}, \tilde{\Theta }_1^i(\varvec{\xi }))- \chi _2(\xi _i)\rho _N(\mathbf {x}, \varvec{\xi }) \right) \\&=(\Delta _N + F^*_N)\rho _N - \alpha _1 N \rho _N + \alpha _2 \sum _{i=1}^N \rho _N(\mathbf {x}, \tilde{\Theta }_1^i(\varvec{\xi })). \end{aligned}$$

Thus, by applying Feynman–Kac’s formula and continuity argument, we get

$$\begin{aligned} \inf _{(\mathbf {x},\varvec{\xi }) \in \Pi ^N}\rho _N(t) > 0. \end{aligned}$$

Indeed, let us consider a d-dimensional process \({\bar{X}}^i_t\) satisfying

$$\begin{aligned} {\bar{X}}^i_t = {\bar{X}}^i_0 +\sigma (1)\int _0^t \,dB^i_s + \frac{1}{N} \sum _{j=1}^N \int _0^t F({\bar{X}}^i_s-{\bar{X}}^j_s, 1)\,ds, \quad i=1,\dots , N, \end{aligned}$$

where \(\bar{\mathbf {X}}_0^N = ({\bar{X}}^1_0, \dots , {\bar{X}}^N_0) = \mathbf {x}\). Then, the generator of the process \(\bar{\mathbf {X}}_t^N = ({\bar{X}}^1_t, \dots , {\bar{X}}^N_t)\) is the following differential operator:

$$\begin{aligned} {\mathcal {G}}_N := \frac{\sigma (1)^2}{2} \sum _{i=1}^N \Delta _{x_i} -\frac{1}{N} \sum _{i,j=1}^N F(x_i-x_j, 1) \cdot \nabla _{x_i}. \end{aligned}$$

Since \(x \mapsto F(x,1)\) is globally Lipschitz and \(F \in L^\infty (\Pi )\), we find

$$\begin{aligned} \begin{aligned}&\rho _N(\mathbf {x},1,\dots , 1,t) \\&\quad = \mathbb {E} \exp \left( -\alpha _1 Nt -\int _0^t f(\bar{X}_s)\,ds \right) \rho _N(\bar{\mathbf {X}}_t^N, 1, \dots , 1, 0) \\&\qquad + \alpha _2 \sum _{k=1}^N \mathbb {E} \int _0^t \exp \left( -\alpha _1 N(t-s)-\int _s^t \sum _{i,j=1}^N f({\bar{X}}_\tau )\,d\tau \right) \rho _N(\bar{\mathbf {X}}_s^N, {\tilde{\Theta }}_1^k(\varvec{\xi }),s)\,ds \end{aligned} \end{aligned}$$
(B.5)

satisfies

$$\begin{aligned} \partial _t \rho _N(\mathbf {x},\varvec{\xi }) = {\mathcal {G}}_N \rho _N(\mathbf {x},\varvec{\xi }) - \left( N + f(x) \right) \rho _N(\mathbf {x},\varvec{\xi }) + \alpha _2 \sum _{i=1}^N \rho _N(\mathbf {x}, {\tilde{\Theta }}_1^i(\varvec{\xi })). \end{aligned}$$

Here

$$\begin{aligned} f(x) := \frac{1}{N} \sum _{i,j=1}^N (\nabla _{x_i} \cdot F)(x_i-x_j, 1) \end{aligned}$$

and

$$\begin{aligned} f({\bar{X}}_t) := \frac{1}{N} \sum _{i,j=1}^N (\nabla _{x_i} \cdot F)({\bar{X}}^i_t-{\bar{X}}^j_t, 1). \end{aligned}$$

On the other hand, for fixed \(j \in \{1,\dots , N\}\), noticing

$$\begin{aligned} \chi _1({\tilde{\Theta }}_1^i(\varvec{\xi })_j) = 1 - \delta _{i}, \quad \chi _2({\tilde{\Theta }}_1^i(\varvec{\xi })_j) = \delta _{ij}, \end{aligned}$$

and

$$\begin{aligned} {\tilde{\Theta }}_j^i({\tilde{\Theta }}_1^i(\varvec{\xi })) = {\tilde{\Theta }}_j^i(\varvec{\xi }) \quad \text{ for } i=1,2, \end{aligned}$$

we obtain

$$\begin{aligned}&\partial _t \rho _N(\mathbf {x}, {\tilde{\Theta }}_1^i(\varvec{\xi })) \\&\quad = (\Delta _N + F^*_N)\rho _N(\mathbf {x}, {\tilde{\Theta }}_1^i(\varvec{\xi }))\\&\qquad + \alpha _1 \sum _{j=1}^N \left( \delta _{ij}\rho _N(\mathbf {x}, \tilde{\Theta }_2^i(\varvec{\xi })) - (1-\delta _{ij})\rho _N(\mathbf {x}, \tilde{\Theta }_1^i(\varvec{\xi })) \right) \\&\quad + \alpha _2 \sum _{j=1}^N \left( (1 - \delta _{ij}) \rho _N(\mathbf {x}, {\tilde{\Theta }}_1^i(\varvec{\xi })) - \delta _{ij} \rho _N(\mathbf {x}, \tilde{\Theta }_1^i(\varvec{\xi })) \right) . \end{aligned}$$

Since \(\rho _N(\mathbf {x}, {\tilde{\Theta }}_2^i(\varvec{\xi })) = \rho _N(\mathbf {x}, 1,\dots , 1) = \rho _N(\mathbf {x}, \varvec{\xi })\), we find

$$\begin{aligned} \begin{aligned}&\partial _t \rho _N(\mathbf {x}, {\tilde{\Theta }}_1^i(\varvec{\xi })) \\&\quad = (\Delta _N + F^*_N)\rho _N(\mathbf {x}, {\tilde{\Theta }}_1^i(\varvec{\xi })) \\&\qquad + \alpha _1\rho _N(\mathbf {x}, \varvec{\xi }) - \left( \alpha _1(N-1) - \alpha _2 (N-2) \right) \rho _N(\mathbf {x}, {\tilde{\Theta }}_1^i(\varvec{\xi })). \end{aligned} \end{aligned}$$
(B.6)

We then define

$$\begin{aligned} { {\mathcal {T}} }:= \left\{ t> 0 : \inf _{\mathbf {x}\in \mathbb {T}^{dN}} \rho _N(\mathbf {x}, 1, \dots , 1, s) > 0 \quad \text{ for } \text{ all } \quad s \in [0,t)\right\} . \end{aligned}$$

By the continuity of \(\rho _N\), it is clear that \({ {\mathcal {T}} }\ne \emptyset \). Set \(t_* := \sup { {\mathcal {T}} }\) and suppose \(t_* < \infty \). Then by definition, \(\lim _{t \rightarrow t_*-}\inf _{\mathbf {x}\in \mathbb {T}^{dN}}\rho _N(\mathbf {x}, 1, \dots , 1, t) =0\). On the other hand, applying Feynman–Kac’s formula to (B.6) and using the positivity assumption on \(\inf _{(\mathbf {x},\varvec{\xi }) \in \Pi ^N}\rho _N(0)\) yield

$$\begin{aligned} \min _{i=1,\dots ,N}\inf _{\mathbf {x}\in \mathbb {T}^{dN}}\rho _N(\mathbf {x}, \tilde{\Theta }_1^i(\varvec{\xi }),t) \ge 0 \end{aligned}$$

for \(t \in [0,t_*)\) due to \(\rho _N(\mathbf {x}, 1,\dots ,1,t) > 0\) for \(t \in [0,t_*)\). We then combine this with (B.5) to get

$$\begin{aligned} \rho _N(\mathbf {x},1,\dots , 1,t) \ge \mathbb {E} \exp \left( -\alpha _1 Nt -\int _0^t f({\bar{X}}_s)\,ds \right) \rho _N(\bar{\mathbf {X}}_t^N, 1, \dots , 1, 0). \end{aligned}$$

Taking the infimum over \(x \in \mathbb {T}^{dN}\) and limit \(t \rightarrow t_*-\) on both sides of the above gives

$$\begin{aligned} 0 = \lim _{t \rightarrow t_*-}\inf _{\mathbf {x}\in \mathbb {T}^{dN}}\rho _N(\mathbf {x}, 1, \dots , 1, t) > 0, \end{aligned}$$

and this is a contradiction. Hence, \(t_* = \infty \).

Now we assume that \(\rho _N\) is positive for \(m < N\) and consider the case \(m+1\). Note that the particles are indistinguishable (see Remark 3). Thus, without loss of generality, we may assume that \(\xi _i = 2\) for \(i=1,\dots ,m+1\), and \(\xi _i =1\) for \(i=m+2,\dots ,N\). We notice that the following argument also holds when \(m+1 = N\). Denoting by

$$\begin{aligned} u_N^{m+1} := \rho _N(\mathbf {x}, \underbrace{2,\dots ,2}_{m+1},\underbrace{1,\dots ,1}_{N-m-1}) =: \rho _N(\mathbf {x}, \varvec{\xi }_0), \end{aligned}$$

we get

$$\begin{aligned} \partial _t u_N^{m+1}&= (\Delta _N + F^*_N)u_N^{m+1} + \alpha _1 \sum _{i=1}^{m+1} \rho _N(\mathbf {x}, {\tilde{\Theta }}^i(\varvec{\xi }_0)) \\&\quad - \left( \alpha _1(N-m+1) + \alpha _2 (m+1)\right) u_N^{m+1}+ \alpha _2 \sum _{m+2}^N \rho _N(\mathbf {x}, {\tilde{\Theta }}^i_1(\varvec{\xi }_0)). \end{aligned}$$

On the other hand, \({\tilde{\Theta }}^i(\varvec{\xi }_0) = (2,\dots , 2,1,2,\dots ,2,1,\dots ,1)\), i.e., it has only m 2’s, thus by the inductive hypothesis, \(\rho _N(\mathbf {x}, {\tilde{\Theta }}^i(\varvec{\xi }_0))\) is positive for \(i=1,\dots , m+1\). Hence,

$$\begin{aligned} \partial _t u_N^{m+1}&\ge (\Delta _N + F^*_N)u_N^{m+1} - \left( \alpha _1(N-m+1) + \alpha _2 (m+1)\right) u_N^{m+1}\\&\quad + \alpha _2 \sum _{m+2}^N \rho _N(\mathbf {x}, \tilde{\Theta }^i_1(\varvec{\xi }_0)) . \end{aligned}$$

Similarly as before, we next observe the equation for \(\rho _N(\mathbf {x}, {\tilde{\Theta }}^i_1(\varvec{\xi }_0))\), and then again apply Feynman–Kac’s formula and continuity argument to the above to have \(u_N^{m+1} (t) > 0\) under the assumption that \(\inf _{(\mathbf {x},\varvec{\xi }) \in \Pi ^N}\rho _N(0) > 0\). Then by the inductive reasoning, we conclude \(\rho _N(t) > 0 \). \(\square \)

We have finished the proof of (3.5).

Rights and permissions

Springer Nature or its licensor holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Ahn, J., Chae, M., Choi, YP. et al. Propagation of Chaos in the Nonlocal Adhesion Models for Two Cancer Cell Phenotypes. J Nonlinear Sci 32, 92 (2022). https://doi.org/10.1007/s00332-022-09854-1

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s00332-022-09854-1

Keywords

Mathematics Subject Classification

Navigation