17
Journal of Theoretical Biology 239 (2006) 334–350 A cellular automata model of tumor–immune system interactions D.G. Mallet a,b, , L.G. De Pillis b a School of Mathematical Sciences, Queensland University of Technology, GPO Box 2434, Brisbane, QLD 4001, Australia b Center for Quantitative Life Sciences and Department of Mathematics, Harvey Mudd College, 301 East Twelfth Street, Claremont, CA 91711, USA Received 24 November 2004; received in revised form 21 July 2005; accepted 1 August 2005 Available online 15 September 2005 Abstract We present a hybrid cellular automata–partial differential equation model of moderate complexity to describe the interactions between a growing tumor next to a nutrient source and the immune system of the host organism. The model allows both temporal and two-dimensional spatial evolution of the system under investigation and is comprised of biological cell metabolism rules derived from both the experimental and mathematical modeling literature. We present numerical simulations that display behaviors which are qualitatively similar to those exhibited in tumor–immune system interaction experiments. These include spherical tumor growth, stable and unstable oscillatory tumor growth, satellitosis and tumor infiltration by immune cells. Finally, the relationship between these different growth regimes and key system parameters is discussed. r 2005 Elsevier Ltd. All rights reserved. Keywords: Tumor; Immune; Cellular automata; Lymphocyte infiltration; Oscillatory growth 1. Introduction The response of the immune system to mutated and potentially cancerous cells is the body’s natural defense against tumor growth. Both the functioning of the immune system and the growth of tumors involve highly complex processes, and coupled, the tumor–immune interactions form an elaborate system that is not yet fully understood by either experimentalists or theoreti- cians. Tumor growth and the dynamics of the immune system have been a significant focus for mathematical modeling over the past three decades. Evolving from the early chemical diffusion and differential equation models of Burton (1966) and Greenspan (1972), descriptions of tumor growth have been presented more recently using partial differential equations (PDEs) (see for example Byrne and Chaplain, 1996, 1997; Owen and Sherratt, 1997; Mallet, 2004; Pettet et al., 2001) and cellular automata (CA) (see for example Alarcon et al., 2003; Dormann and Deutsch, 2002; Bard Ermentrout and Edelstein-Keshet, 1993; Ferreira et al., 2002; Kansal et al., 2000; Patel et al., 2001). An excellent review of mathematical models of tumor growth was recently given by Araujo and McElwain (2004). Similarly, a variety of mathematical models exist describing the immune system in its various roles. Perelson provides a substantial review of the immune system before demonstrating the process of modeling the immune system from the point of view of a physicist (Perelson and Weisbuch, 1997). Seiden and Celada (1992) developed a computer program based on a bit- string cellular automata called the IMMSIM model. The IMMSIM model in its current form includes both the humoral and cellular immune responses. A number of researchers have also modeled the kinetics of the cellular immune response, including Merrill (1982), Callewaert et al. (1988), and Perelson and Mackean (1984). A number of mathematical models also have coupled tumor growth with immune system dynamics. Usually, these models are fully deterministic, comprised of a ARTICLE IN PRESS www.elsevier.com/locate/yjtbi 0022-5193/$ - see front matter r 2005 Elsevier Ltd. All rights reserved. doi:10.1016/j.jtbi.2005.08.002 Corresponding author. School of Mathematical Sciences, Queens- land University of Technology, GPO Box 2434, Brisbane, QLD 4001, Australia. Tel.: +617 3864 2354; fax: +617 3864 2310. E-mail address: [email protected] (D.G. Mallet).

A cellular automata model of tumor–immune system interactionszapata/Journalclub/Mallet_2005.pdf · Journal of Theoretical Biology 239 (2006) 334–350 A cellular automata model

  • Upload
    others

  • View
    1

  • Download
    0

Embed Size (px)

Citation preview

Page 1: A cellular automata model of tumor–immune system interactionszapata/Journalclub/Mallet_2005.pdf · Journal of Theoretical Biology 239 (2006) 334–350 A cellular automata model

Journal of Theoretical Biology 239 (2006) 334–350

A cellular automata model of tumor–immune system interactions

D.G. Malleta,b,!, L.G. De Pillisb

aSchool of Mathematical Sciences, Queensland University of Technology, GPO Box 2434, Brisbane, QLD 4001, AustraliabCenter for Quantitative Life Sciences and Department of Mathematics, Harvey Mudd College, 301 East Twelfth Street, Claremont, CA 91711, USA

Received 24 November 2004; received in revised form 21 July 2005; accepted 1 August 2005Available online 15 September 2005

Abstract

We present a hybrid cellular automata–partial differential equation model of moderate complexity to describe the interactionsbetween a growing tumor next to a nutrient source and the immune system of the host organism. The model allows both temporaland two-dimensional spatial evolution of the system under investigation and is comprised of biological cell metabolism rules derivedfrom both the experimental and mathematical modeling literature. We present numerical simulations that display behaviors whichare qualitatively similar to those exhibited in tumor–immune system interaction experiments. These include spherical tumor growth,stable and unstable oscillatory tumor growth, satellitosis and tumor infiltration by immune cells. Finally, the relationship betweenthese different growth regimes and key system parameters is discussed.r 2005 Elsevier Ltd. All rights reserved.

Keywords: Tumor; Immune; Cellular automata; Lymphocyte infiltration; Oscillatory growth

1. Introduction

The response of the immune system to mutated andpotentially cancerous cells is the body’s natural defenseagainst tumor growth. Both the functioning of theimmune system and the growth of tumors involve highlycomplex processes, and coupled, the tumor–immuneinteractions form an elaborate system that is not yetfully understood by either experimentalists or theoreti-cians.

Tumor growth and the dynamics of the immunesystem have been a significant focus for mathematicalmodeling over the past three decades. Evolving from theearly chemical diffusion and differential equationmodels of Burton (1966) and Greenspan (1972),descriptions of tumor growth have been presented morerecently using partial differential equations (PDEs) (seefor example Byrne and Chaplain, 1996, 1997; Owen and

Sherratt, 1997; Mallet, 2004; Pettet et al., 2001) andcellular automata (CA) (see for example Alarcon et al.,2003; Dormann and Deutsch, 2002; Bard Ermentroutand Edelstein-Keshet, 1993; Ferreira et al., 2002; Kansalet al., 2000; Patel et al., 2001). An excellent review ofmathematical models of tumor growth was recentlygiven by Araujo and McElwain (2004).

Similarly, a variety of mathematical models existdescribing the immune system in its various roles.Perelson provides a substantial review of the immunesystem before demonstrating the process of modelingthe immune system from the point of view of a physicist(Perelson and Weisbuch, 1997). Seiden and Celada(1992) developed a computer program based on a bit-string cellular automata called the IMMSIM model. TheIMMSIM model in its current form includes both thehumoral and cellular immune responses. A number ofresearchers have also modeled the kinetics of the cellularimmune response, including Merrill (1982), Callewaertet al. (1988), and Perelson and Mackean (1984).

A number of mathematical models also have coupledtumor growth with immune system dynamics. Usually,these models are fully deterministic, comprised of a

ARTICLE IN PRESS

www.elsevier.com/locate/yjtbi

0022-5193/$ - see front matter r 2005 Elsevier Ltd. All rights reserved.doi:10.1016/j.jtbi.2005.08.002

!Corresponding author. School of Mathematical Sciences, Queens-land University of Technology, GPO Box 2434, Brisbane, QLD 4001,Australia. Tel.: +617 3864 2354; fax: +617 3864 2310.

E-mail address: [email protected] (D.G. Mallet).

Page 2: A cellular automata model of tumor–immune system interactionszapata/Journalclub/Mallet_2005.pdf · Journal of Theoretical Biology 239 (2006) 334–350 A cellular automata model

series of ordinary or partial differential equations(ODEs or PDEs) describing the dynamics of, forexample, tumor cells, host cells, and immune cells(Arciero et al., 2004; Bellomo et al., 1999; Bellomoand Preziosi, 2000; Galach, 2003; Kirschner andPanetta, 1998; Kuznetsov and Knott, 2001; Lin, 2004;Matzavinos and Chaplain, 2004; Owen and Sherratt,1997, 1998; Owen et al., 2004; de Pillis and Radunskaya,2003a; Takayanagi and Ohuchi, 2001). An excellentreview of tumor–immune system models, both at thecellular and sub-cellular levels, is given by Adam andBellomo (1997). Immunotherapy has also been investi-gated using similar deterministic models coupled withoptimal control theory (Burden et al., 2004; de Pillis andRadunskaya, 2001, 2003b).

Most of these tumor–immune system models are fullydeterministic, and although there are several excellentmodels that include spatial interactions between tumorand immune cells using chemotaxis terms in thegoverning PDEs, simulation results often focus mainlyon one-dimensional or radially symmetric spatialgrowth and the temporal evolution of the cell speciesunder investigation (Matzavinos and Chaplain, 2004;Owen and Sherratt, 1997; Owen et al., 2004). In thispaper, we also model temporal evolutions but further-more include two-dimensional spatial evolution byemploying a hybrid cellular automata–partial differen-tial equation modeling approach.

This deterministic–stochastic approach has the bene-fits of being conceptually accessible and computation-ally straightforward to implement. Unlike the spatio-temporal models noted above, this model allows for theconsideration of individual cell behavior and associatedrandomness, rather than applying a general rule to acollection of cells (for example, with regard to cellmigration). Because of the structure of the model, whichsimulates chemical diffusion through deterministicPDEs and cell behavior through a set probabilisticrules, the model and its computational implementationare very flexible. It can easily be extended or modified asnew data become available or when different situationsarise requiring the inclusion of new chemical species orcell behavior rules.

Few if any other models, consider the spatio-temporaland partially stochastic interactions between individualtumor cells and multiple populations of individuallyrecognised immune cells, such as is the case in this work.Furthermore, biologists and immunologists also have alimited understanding of such interactions. As such webelieve it is important to focus on early tumor growthbefore tackling the more complicated vascularised stagesof growth in such an investigation.

It should also be noted that the strategy employed inthis work to describe the tumor dynamics and interac-tions with the immune system is quite different from thecontinuous–discrete hybrid used by Anderson and

Chaplain (1998) in their models of angiogenesis.Anderson and Chaplain employ discrete models closelyrelated to continuous PDE-based angiogenesis modelswhereas in this work we consider a mixture ofcontinuous PDEs for chemical quantities and a discretecell-based description for biological cell species withphenomenologically sourced probabilities for cell dy-namics. It is also different from the off-lattice approach,another technique that can be used in this type ofresearch, which is employed by Drasdo and coworkers(Drasdo and Loeffler, 2001; Galle et al., 2005) andallows for the modeling of forces on the cells (such asdeformation forces).

The hybrid CA–PDE modeling approach has beensuccessfully used in the past to model tumor growth,chemotherapeutic treatment and the effects of vascular-ization on tumor growth (Patel et al., 2001; Ferreira etal., 2002; Alarcon et al., 2003; Ferreira et al., 2003).Here we use reaction–diffusion PDEs to describechemical species such as growth nutrients, and a cellularautomata strategy to track the tumor as well as twodistinct immune cell species. Together, these elementssimulate the growth of the tumor and the interactions ofthe immune cells with the tumor–growth.

Using the hybrid CA–PDE model, we aim todemonstrate the combined effects of the innate andspecific immune systems on the growth of a two-dimensional tumor. This is accomplished through thedevelopment of a model with cellular behaviors relatedto those described in the experimental literature and byconsidering the ODE models of tumor–immune systeminteractions developed in other works (such as Kuznet-sov and Knott, 2001; de Pillis and Radunskaya, 2001,2003b).

This model is useful because it allows for predictionsto be made with regard to the behavior of the initialstages of solid tumor growth and the growth ofundetectable metastases. In this model, we study acluster of tumor cells to which nutrient is made availablethrough a nearby blood vessel. Note that in a model oflater stage development, one might wish to incorporatenutrient delivery through blood vessel sources inter-spersed throughout the tumor cell cluster as has beenattempted by Alarcon et al. (2003), McDougal et al.(2002) and Gatenby and Gawlinski (1996). However, inthis work we focus on the early stage model as the studyof early stage tumor models continues to be of interest(see, for example, Owen and Sherratt, 1998, 1999;Smallbone et al., 2005; Byrne and Chaplain, 1995;Byrne, 1997; Please et al., 1999; Ward and King, 1997,1999; Franks et al., 2003, 2003, 2005; Owen et al., 2004;Preziosi, 2003; Xu, 2004). One of the reasons formodeling tumor growth in its early stages is to allowfor the development and validation of a foundationalmodel prior to the inclusion of potentially confound-ing model features such as full vascularization and

ARTICLE IN PRESSD.G. Mallet, L.G. De Pillis / Journal of Theoretical Biology 239 (2006) 334–350 335

Page 3: A cellular automata model of tumor–immune system interactionszapata/Journalclub/Mallet_2005.pdf · Journal of Theoretical Biology 239 (2006) 334–350 A cellular automata model

metastatic behavior. Another reason is to allow for themodeling of the situation in which a patient whoseprimary tumor has been resected, but who mayadditionally require post-operative therapy to ensurethat undetectable metastases are being treated. Theseundetectable metastases are reasonably modeled as earlystage tumors.

We will demonstrate different types of tumor growthincluding spherical and papillary (branchy) growth,stable and unstable oscillatory growth, and lympho-cyte-infiltrated growth, and show the dependence ofthese different morphologies on key model parametersrelated to the tumor–immune system interactions. Themodel developed here is also useful as it provides anextension to previous models lacking an immune systemcomponent, that may be built upon in the future withthe addition of other cell and chemical species in orderto test and validate new hypotheses.

In a study of cell transfer therapy for metastaticmelanoma patients, Rosenberg et al. (2003) comment onthe difficulty of deriving meaningful results from humanexperiments due to the variations in cell types, tumortypes, immune states, and more fundamentally thehuman subjects themselves. While Rosenberg et al.suggest a solution to such a problem is to treat the samepatient in differing ways over a period of time, anothermore flexible and less hazardous method is throughmathematical modeling—the pathway taken in thispaper to examine the effects of the immune system ontumors.

In Section 2 we present a short overview of thebiology underlying the mathematical model that will bedeveloped in Section 3. In Section 4, a number ofrepresentative simulation outputs are presented alongwith a discussion of the model results in general. Finally,in Section 5, the results of the model simulations arediscussed and related to the biological problem, and anumber of ideas for future research are proposed.

2. Biological background

Consider first the biology underlying the growth of atumor. In normal tissue, division and death of cells aretightly balanced. This balance allows the body toconstruct and maintain the tissue so that it may carryout its particular purpose in the survival of theorganism. The growth of a tumor starts with a singlemutated cell that is able to proliferate inappropriatelyby either avoiding apoptosis or undergoing excessiveproliferation. This leads to an imbalance in the celldeath:division ratio when compared with that observedin normal tissue (Kleinsmith et al., 2001; Weinberg,1996).

In this paper we consider a tumor in the early stage ofgrowth. Avascular growth, that which occurs when the

tumor does not have its own vasculature, is the earlyphase of tumor growth that follows cell mutation(Byrne, 1999). During this stage, nutrients required forgrowth (such as glucose, oxygen (Folkman and Hoch-berg, 1973; Sutherland, 1988) or some other relevantnutrient) are supplied to the tumor via diffusion fromdistant blood vessels (Byrne, 1999). Aggressive growth,including the possibility of metastasis, occurs after anavascular tumor has developed its own vascular networkand will not be considered in the model developed in thispaper. Rather, we focus on the early stages of tumorgrowth in which the tumor is simply adjacent to nutrientsupplying vasculature to allow for an investigation ofthe initial interactions between the immune system andthe emerging tumor.

As a result of the external nutrient supply, avasculartumors often develop into approximately radiallysymmetric structures. Ferreira et al. (2002) have alsoshown that branch-like structures develop when tumorcells divide rapidly and consume nutrients required formitosis much more rapidly than those required forsurvival. In radially symmetric tumors, the growingtumor is characterized by up to three layers—an outerrim of dividing cells with easy access to nutrients, aninner shell of non-dividing cells subjected to lowernutrient levels, and a central core of necrotic (dead) cellsthat die due to extremely low nutrient concentrations.

Tumor growth is often suppressed either naturally bythe immune system of the host, or by outside interven-tion using chemotherapeutic and other drug treatments.The purpose of this study is to investigate the effects ofthe immune system on tumor growth. The effects ofchemotherapeutic drugs have been investigated insimilar work by Ferreira et al. (2003) and in the ODEmodels of de Pillis and Radunskaya (2001).

Fundamental to the immune system are the conceptsof ‘‘self ’’ and ‘‘non-self ’’. The components of theimmune system are located throughout the body andare charged with the responsibility of detecting anddestroying entities which are foreign to the body (non-self), such as tumor cells, while simultaneously allowingentities that belong in the body (self) to remain andcarry out their function (Paul, 2003; Perelson andWeisbuch, 1997). Tumor cells present different antigensfrom those of the cells from which they have mutated.These are the antigens that alert the immune system tothe non-self nature of the tumor cell. When there is afailure in the immune system’s surveillance of tumorcells, or when the immune suppression of tumor cells isinsufficient, a tumor is able to grow and invadesurrounding tissue (Kleinsmith et al., 2001).

While other components (such as macrophages) exist,in the model developed in this research we consider onlytwo cell types to comprise the immune system. Inparticular, we model the natural killer (NK) cells of theinnate immune system and the cytotoxic T lymphocytes

ARTICLE IN PRESSD.G. Mallet, L.G. De Pillis / Journal of Theoretical Biology 239 (2006) 334–350336

Page 4: A cellular automata model of tumor–immune system interactionszapata/Journalclub/Mallet_2005.pdf · Journal of Theoretical Biology 239 (2006) 334–350 A cellular automata model

(CTLs) of the specific immune system. Essentially, it isin this way that the model developed herein is differentfrom that of Ferreira et al. Both are hybrid PDE–CAmodels, but this work introduces the impact of theimmune response on tumor development.

Natural killer cells are able to find and eliminatemalignant cells while they undertake surveillance of thebody for foreign components (Kleinsmith et al., 2001).Forming part of the innate immune system, naturalkiller cells are cytotoxic cells that are highly effective inlysing multiple (but specific) tumor cell lines (Paul,2003). Unlike cells of the specific immune system, whichare drawn to a location due to the presence of antigen,the natural killer cells are constantly present guardingthe body from infection and disease.

On the other hand, CTLs are able either to lyse or toinduce apoptosis in cells presenting specific antigens,such as tumor cells (Paul, 2003). Unlike NK cells, CTLsare only able to recognize a specific antigen or tumor cellline. It is known however, that these cells are able todestroy more than one tumor cell during their life cyclewhile a single natural killer cell generally kills very few(Kuznetsov, 1997). After destroying the target cell, theCTLs move on in search of other antigen presentingcells.

Using the modeling framework described in thispaper, we also demonstrate lymphocyte infiltration oftumors. This is of particular interest given recentexperimental research that suggests improved survivalrates for patients with intratumoral immune cells(Zhang et al., 2003). Infiltration of T cells into thetumor mass can also lead to fibrosis and necrosis, andtherefore tumor destruction and antitumor immunity(Schmollinger et al., 2003; Soiffer et al., 1998). Tumorsreduced in size are then more amenable to treatment bytraditional radiation methods. Numerical solutionsproduced using the model developed in this paper arein qualitative agreement with the experimental resultsdemonstrated by Zhang et al. (2003), Schmollinger et al.(2003), and Soiffer et al. (1998).

In the following section, the above biological descrip-tion will be translated into the mathematical model usedto simulate tumor growth and the interaction of thetumor with the immune system.

3. Model formulation

In this model we consider the nutrient limited growthof an early stage tumor and the dynamic interplaybetween the immune system and the growing tumor.The model involves a combination of reaction–diffusionequations for the nutrient species and numerous rules ofevolution for the cellular automaton description of thevarious cell types that comprise the tissue–tumorenvironment.

We consider a tumor growing on a square domainO ! "0;L# $ "0;L#, where L is the length of eachside of the domain. The spatial domain represents atwo-dimensional patch of tissue that is suppliedwith nutrients by blood vessels that occupy the y ! 0and y ! L boundaries, as shown in Fig. 1. Theremainder of the space is partitioned into a regular gridin which the various cell types reside, and through whichthe nutrient species diffuse. The grid is partitioned insuch a way that each cellular automata grid elementcorresponds in size with the actual biological cells ofinterest (10–20 mm; Alarcon et al., 2003; Lin, 2004).While finer grids can be used for the nutrient PDEsolver, sufficient accuracy was gained with the matchinggrids and this allowed for faster computation ofsolutions.

We simulate the temporal progression of thesystem using two main steps. First, the partial differ-ential equations for the nutrient species are solved, thenwith dependence upon the new nutrient fields thecellular activities (such as motion and division) arecarried out. Each temporal iteration therefore corre-sponds to the period of tumor cell division. That is, aperiod of approximately 0.5–10 days, depending on thecell type in question (Kirschner and Panetta, 1998;Riedel, 2004). The cellular automaton rules, along witha description of the reaction–diffusion equations aregiven below. For reference, the variables and parametersused in the model are listed along with descriptions inTable 1.

ARTICLE IN PRESS

FIXED SOURCE

FIXED SOURCE

WRAPPED

WRAPPED

y

X

Fig. 1. Schematic of the cellular automata physical domain. Theconditions imposed on the four boundaries are indicated. The solidbars (top and bottom) represent the capillaries while different cell typesare shown filling the spaces in the grid.

D.G. Mallet, L.G. De Pillis / Journal of Theoretical Biology 239 (2006) 334–350 337

Page 5: A cellular automata model of tumor–immune system interactionszapata/Journalclub/Mallet_2005.pdf · Journal of Theoretical Biology 239 (2006) 334–350 A cellular automata model

3.1. Reaction– diffusion equations

Reaction–diffusion equations have been widely usedsince the early models of (Greenspan, 1972) and Burton(1966) for the mathematical treatment of nutrientspecies, and many other chemical species, in tumor-growth models. In a similar manner, reaction–diffusionpartial differential equations are also employed in thiswork to describe the distribution of two chemicalsnecessary for mitosis and cell survival.

As in the work of Ferreira et al. (2002, 2003), here weconsider two nutrient species—the first nutrient being anecessary component of the cell division processes, whilethe second is essential for cellular survival. The nutrientsdiffuse throughout the tissue space, and as they do sothey are consumed by the different cells that are residentin tissue.

The nutrient species are governed by the followingequations:

qN

qt! DNr2N " k1HN " k2TN " k3IN, (1)

qM

qt! DMr2M " k4HM " k5TM " k6IM, (2)

where N and M represent the proliferation nutrient andsurvival nutrient, respectively. The cell species areidentified by H for host cells (normal tissue), T fortumor cells, and I for immune cells. Also, DN and DM

are the diffusion coefficients for the two nutrients, k1; k2

and k3 are the rates of consumption of proliferationnutrient by host cells, tumor cells and immune cells,respectively, while k4; k5 and k6 are the correspondingrates for the survival nutrient.

So as to maintain continuity of comparison with thework of Ferreira et al., upon which the above equationsare based, we assume that the diffusion coefficients forboth nutrients are equal. That is, DM ! DN ! D.Furthermore, we make similar assumptions for theconsumption rates of the two chemicals by non-tumorous cells, and consider both normal tissue andimmune cells to consume nutrients at the same rate suchthat k1 ! k4 ! k3 ! k6. To model the increased nutrientconsumption of tumor cells compared with that ofnormal cells, we further assume that the consumptionrate of the tumor cells is a constant multiple of the hostcell consumption rates. That is, k2 ! lNk1 andk5 ! lMk1, where lNX1 and lMX1 determine theexcess consumption by the tumor cells of the two typesof nutrient.

As mentioned above, we assume blood vessels passthrough the area of interest and take for example, a pairof blood vessels residing along the boundaries y ! 0 andy ! L. Given that the blood vessel is the primary sourceof nutrient supply, the conditions at those boundariesfor both reaction–diffusion equations take on the formN#x; 0$ ! N#x;L$ ! N0 and M#x; 0$ !M#x;L$ !M0.On the sides (x ! 0 and x ! L) we assume periodicboundary conditions.

To non-dimensionalise the reaction–diffusion equa-tions, we assume the same variable transformations usedby Ferreira et al. (2002), such that

t !Dnt

L2; #x; y$ !

nx

L;ny

L

! "

N !N

N0; M !

M

M0, #3$

where variables under hats are dimensionless. Withthese transformations, Eqs. (1) and (2) in their non-dimensionalised form become

qN

qt! r2N " a2#H % I$N " lNa2TN, (4)

qM

qt! r2M " a2#H % I$M " lMa2TM, (5)

where a2 ! k1L2=Dn2 is the dimensionless rate of

consumption of nutrient by host and immune cells andhats have been dropped for notational simplicity.

ARTICLE IN PRESS

Table 1Variables and model parameters used in the hybrid model

Variable Description

N#x; y; t$ mitosis nutrient concentrationM#x; y; t$ survival nutrient concentrationH host cell numberT tumor cell numberI total immune cell number

Parameter Description

a rate of consumption of nutrient by host and immunecells

lN excess consumption factor of tumor over non-tumorcells of N nutrient

lM excess consumption factor of tumor over non-tumorcells of M nutrient

Pnec probability of tumor cell death due to necrosisynec shape parameter for necrotic death probability curvePimdth probability of tumor cell death due to the immune

systemPdiv probability of tumor cell divisionydiv shape parameter for cell division probability curvePmig probability of tumor cell migrationymig shape parameter for migration probability curvek number of possible tumor cell kills for a single

CD8+ cellI0 background level of natural killer cellsPnk probability of the production of a new natural killer

cellPL probability of induction of a CTL due to CTL/TC

interactionyL shape parameter for CTL induction probability

curvePLD probability of CTL deathyLD shape parameter for CTL death probability curve

D.G. Mallet, L.G. De Pillis / Journal of Theoretical Biology 239 (2006) 334–350338

Page 6: A cellular automata model of tumor–immune system interactionszapata/Journalclub/Mallet_2005.pdf · Journal of Theoretical Biology 239 (2006) 334–350 A cellular automata model

The change to non-dimensional variables also altersthe source conditions on the y ! 0 and y ! L bound-aries. These become N"x; 0# ! N"x; n# ! 1 andM"x; 0# !M"x; n# ! 1.

3.2. Cellular automata rules

While the chemical species are governed by determi-nistic reaction–diffusion equations, the evolution of thefour cell species considered in this model proceedsaccording to a combination of probabilistic and directrules. These are outlined below. As in Ferreira et al., weconsider a simplification of the host cells being passivewhere, other than their consumption of nutrients, theyallow tumor cells to freely divide and migrate.

3.2.1. Tumor cellsThe tumor cells in this model undergo the processes of

division, death, and movement. These processes dependupon nutrient levels, the presence of cells of the immunesystem, and crowding due to the presence of othertumor cells. At each time step, each cell is randomlyassigned the option to divide, move or die, after whichvarious conditions are checked to determine whether ornot the action will be carried out. While it is possiblethat the time-scales of such processes may vary (betweenor within cell types), for generality we consider the time-scales to coincide with the time step of the numericalsolution method. To effect the various changes to thetumor cell population, we impose the following cellularautomata rules.

Due to their inherent mutations by which they avoidapoptosis, we do not consider natural death of tumorcells. We do however, consider two pathways to tumorcell death:

Death due to insufficient nutrient. If a cell is markedfor death, the probability of cell death,

Pnec ! exp $M

Tynec

! "2" #

(6)

is calculated to determine whether or not the action willbe carried out. This probability term implies thatthe possibility of nutrient related death decreasesas the ratio of nutrient M to the number of tumorcells T increases, with shape parameter ynec. Thisprobability is of the same form as that used by Ferreiraet al. (2002).

Death due to the immune system. When the cells of theimmune system arrive in a close neighborhood of atumor cell, the immune cells try to kill the tumor cell,with an intensity directly related to the strength of thelocal immune system. The probability of death in this

case is given by

Pimdth ! 1$ exp $X

j2ZI j

!22

4

3

5, (7)

where the summation counts the total number ofimmune cells (both NK and CTLs), I, in the neighbor-hood, Z, of the relevant tumor cell. Without furtherexperimental information, we consider Z to include theeight CA elements surrounding element j.

Tumor cells can undergo mitosis provided the locallevel of the mitosis nutrient, N, is sufficient.

Tumor cell division. If a cell is marked for division, theprobability of cell division,

Pdiv ! 1$ exp $N

Tydiv

! "2" #

(8)

is calculated to determine whether mitosis occurs. Thisterm implies that the chance of division increases withthe ratio of nutrient concentration to the number oftumor cells, with a shape parameter ydiv, and is of theform used by Ferreira et al. (2002).

The grid location upon which the daughter cell isplaced depends upon the cells occupying the neighbor-hood of the mother cell. For example, a dividing cellwith at least one host cell or necrotic space surroundingit will place its daughter cell randomly in one of thosenon-cancerous locations and either destroy the host cellor simply replace the necrotic material. On the otherhand, if all elements around the dividing cell are filledwith tumor cells, the daughter cell will be placed in theneighboring element containing the fewest tumor cells.This may be viewed as one approach to modeling tumorcell crowding.

Finally, tumor cells may also be chosen to move to aneighboring location on the cellular automata grid.

Tumor cell migration. Cells marked for migration doso with a probability given by

Pmig ! 1$ exp $TM

ymig

! "2" #

, (9)

such that the probability that the cell migrates increaseswith the tumor cell count of the current element. This isanother way that cell crowding effects are accounted forin the model. This probability term, proposed byFerreira et al. (2002), suggests that the likelihood oftumor cell movement increases with higher levels of thenutrient, M, possibly because nutrients allow the cell toproceed with migratory processes. It has also beenconsidered that a cell may prefer to move away fromareas of low nutrient concentration, and in this case theM in the probability term would appear in thedenominator. Preliminary experiments with this modelalteration have produced little qualitative difference to

ARTICLE IN PRESSD.G. Mallet, L.G. De Pillis / Journal of Theoretical Biology 239 (2006) 334–350 339

Page 7: A cellular automata model of tumor–immune system interactionszapata/Journalclub/Mallet_2005.pdf · Journal of Theoretical Biology 239 (2006) 334–350 A cellular automata model

solutions, however this will be investigated further infuture research.

The grid location to which the migrating cell relocatesagain depends upon the cells that occupy its neighbor-hood. Migrating cells that have necrotic space or hostcells in their immediate neighborhood will randomlymove to one of these elements, replacing the originaloccupant as was the case for a daughter cell above.When all neighboring automaton elements are occupiedby tumor cells, the migrating cell will move to theelement with the fewest cancer cells—again in anattempt to move away from crowded areas of thetumor. While it has not been considered in this work,directed motion such as chemotaxis could be introducedto the model here by requiring the moving cell to movetoward areas of higher nutrient levels.

While Eqs. (6), (8) and (9) are taken from Ferreiraet al. (2002), and Eq. (7) is constructed from a similarpoint of view, different functional forms mayindeed lead to slightly different results from thequantitative viewpoint, and with more experimentaljustification it will be possible to give more accurateprobability functions. This is currently a topic of furtherresearch.

3.2.2. Immune cellsTwo separate immune cell populations are considered

here—the natural killer cells of the innate immunesystem, and cells of the specific immune response,represented by the CTLs.

Here we consider CTLs to be recruited to the tumorlocation from external sources when natural killer cellslyse tumor cells or when CTLs and tumor cells interact.Furthermore, the cells of the specific immune system areable to lyse tumor cells more than once (Kuznetsov,1997). Hence, we assume that CTLs may kill a fixednumber of tumor cells and use k to represent this killparameter.

Like the tumor cells, at each time step, immune cellsthat are near tumor cells are marked with the intent tocarry out tumor cell lysis. On the other hand, thoseimmune cells with no nearby tumor cells are randomlyassigned to either die off or to attempt to migrate insearch of tumor cells.

In this model we assume that natural killer cellnumbers are roughly stable around some ‘‘normal’’constant proportion, I0, of the total number of cells inthe domain. This is achieved by imposing an initialcondition where I0n

2 NK cells are randomly placed overthe domain of interest.

Natural killer cell production. In order to maintainthe normal level of the natural killer cell population,we impose a form of birth on the evolution of theimmune cell population. At each time step and for eachCA grid element, a random number is generatedand compared with the probability of natural killer

cell production:

Pnk ! I0 "1

n2

X

all CA

I . (10)

If the random number is less than Pnk and the element isnot currently occupied by a tumor cell, then a newnatural killer cell will be placed in that element. Thisprobability term, Pnk, effectively compares the currentNK cell population in the domain of interest with theproportion that is expected in the normal situation. Theeffect of this strategy is that it allows NK cells to arrivein open spaces, possibly from below the domain ofcomputation (in an attempt to incorporate at least somethree-dimensionality in the model), but also in freespaces near the two blood vessels (at the top and bottomof domain).

Immune cell lysis of tumor cells. If an immune cellcomes in contact with a tumor cell there is a probabilitythat the immune cell will lyse the tumor cell. This rulecorresponds to the rule for tumor cell death due to theimmune system using Eq. (7).

# If the immune cell is a NK cell it is destroyed alongwith the tumor cell. Furthermore, the CA element isflagged for the induction of the specific immunesystem (through CTLs) at the next time step.# If the immune cell is a CTL, the effectiveness of the

cell is reduced—i.e., the number of remaining tumorcell kills, k, is reduced by one.# If the immune cell is a CTL, the surrounding CA

elements are sampled for further CTL induction. Foreach neighboring free cell a random number iscompared with

PL ! exp "yLP

j2ZTj

0

B@

1

CA

22

64

3

75 (11)

to determine whether or not further induction isinitiated. In the expression for PL the summationcounts the number of tumor cells, T, in theneighborhood, Z, of the relevant CA element, andyL is a parameter determining the shape of theprobability curve. This lymphocyte recruitment facil-itates the satellitosis where T cells have been observedsurrounding tumor cells undergoing apoptosis (Soif-fer et al., 1998). For completeness we note that for aneighboring tumor cell count of zero, PL is defined tobe zero also.

Immune cell death. CTLs are assumed to die by one oftwo means. Firstly, they become functionally useless(and therefore, not tracked by the program) when theircell kill value k is reduced to zero. Secondly, the CTLswill die off or move away from the domain of interest ifthey no longer detect tumor cells in their neighborhood.

ARTICLE IN PRESSD.G. Mallet, L.G. De Pillis / Journal of Theoretical Biology 239 (2006) 334–350340

Page 8: A cellular automata model of tumor–immune system interactionszapata/Journalclub/Mallet_2005.pdf · Journal of Theoretical Biology 239 (2006) 334–350 A cellular automata model

This death occurs with probability

PLD ! 1" exp "yLDP

j2ZTj

0

B@

1

CA

22

64

3

75, (12)

where the summation counts all tumor cells in theneighborhood of the immune cell, yLD is a parameterdetermining the shape of the probability curve, and theprobability of immune cell death is defined to be unitywhen there are no tumor cells nearby.

Immune cell migration. Immune cells that do notencounter a tumor cell at some point in time willcontinue to move randomly about the domain ofinterest.

# If there are tumor cells in neighboring CA elements,the immune cell will choose to attack one of these atrandom.# If there are no tumor cells in neighboring CA

elements, the immune cell will choose an element atrandom and move there.

We point out that while tumor antigenicity is not adirect component of these CA rules, antigenicity levelscan be reflected in the model through the adjustment ofthe immune production and recruitment parameters,such as Pnk; PL and yL.

Together, these cell rules and the partial differentialequations describe the evolution of the tumor–immunesystem environment under consideration. To model anemerging tumor, the initial state of the system is taken asfollows. The cellular automata grid houses a singlemutated cancer cell along with the normal level, I0, ofnatural killer immune cells. The remainder of spaceavailable to cells is occupied by non-tumorous host cells.As in other similar models (such as Alarcon et al., 2003;Ferreira et al., 2002, 2003; Patel et al., 2001), thenutrient species are assumed to be in a temporal steady-state distribution throughout the domain since the time-scale for nutrient diffusion is known to be much shorterthan the time-scale for cell metabolism. The steady-statedistribution being determined during the initial calcula-tion of solutions to Eqs. (4) and (5).

4. Simulations and results

The model described in the previous section wasimplemented using MATLAB Release 13 (The Math-works Inc, 2002). The dimensionless parameter valuesa; lM and lN and shape parameters for tumor cellbehavior have been chosen to correspond with thoseused by Ferreira et al. (2002) for similar growthsimulations. The shape parameters for the probabilities

related to immune cell dynamics are unknown and willbe investigated in this section.

The solution of the model proceeds via a sequence ofsteps and a summary of this algorithm is given below.

(1) Assign parameter values, discretise spatial domainand construct cell data structures with initialconditions.

(2) Solve steady-state PDEs with dependence uponinitial cell conditions, to obtain initial chemicalconcentrations.

(3) Loop over the required number of time steps or untilstop conditions are satisfied:(i) Assign random number to all CA locations for

use in deciding actions of cells.(ii) For each cell location in turn, calculate prob-

abilities of the relevant cell action (that is, tumorcell death, migration, division, or immune celldeath, migration and induction), dependingupon the random number assigned above andthe calculated nutrient concentrations.

(iii) Update the cell data structures according to theprobabilities and random actions decided uponabove.

(iv) Recalculate the nutrient concentrations result-ing from the changed cell conditions.

(v) Stop if tumor reaches domain edge or iseradicated, otherwise proceed to next time stepand return to (i).

The initial conditions are always taken to be a single,mutated cell, so we do not consider sensitivity to theinitial tumor cell distribution. For all results shown anddiscussed in the text to follow, at least 20 simulationshave been carried out in order to give a general result forthe parameter sets reported. Unfortunately the simula-tions are computationally intensive and further investi-gation is therefore quite difficult.

4.1. Immune system-free growth

Our findings confirm those of Ferreira et al. (2002,2003), in which simulations show that tumor morphol-ogy is dictated by relative consumption rates: lowerconsumption rates lead to more compact tumors whilerelatively higher consumption rates lead to a branchiermorphology. Furthermore, we have found that it isindeed possible to simulate tumors that grow firstexponentially, then linearly and finally reach a steady-zone of existence. Unlike the deterministic model ofGreenspan (1972), the stochasticity of this model meansthat small fluctuations in the tumor size are stillobserved—however, it can be claimed that the tumoris in a zone of stability with regard to the total numberof tumor cells in the mass.

ARTICLE IN PRESSD.G. Mallet, L.G. De Pillis / Journal of Theoretical Biology 239 (2006) 334–350 341

Page 9: A cellular automata model of tumor–immune system interactionszapata/Journalclub/Mallet_2005.pdf · Journal of Theoretical Biology 239 (2006) 334–350 A cellular automata model

Fig. 2(a) shows the growth in the total number oftumor cells over time when the tumor is allowed to growin the absence of any immune response. Note theinitially exponential growth phase (iteration 0–200),before a phase of linear growth (iteration 200–800).These growth characteristics mimic the growth rates ofmulticell spheroids described experimentally by Folk-man and Hochberg (1973) and mathematically byGreenspan (1972).

Fig. 2(b) displays the state of the system after 800iterations. A roughly circular tumor with a radius ofabout 200 cells has developed in the center of thedomain and is growing steadily toward the sources ofthe nutrient. Higher tumor cell densities are seen at theperiphery of the tumor where it is surrounded by normalcells comprising the host tissue. In the center of thetumor a necrotic core is beginning to form with somenecrotic material already appearing. The tumor shown isgrowing in a domain that is approximately 10–20mmsquare, and over a time period of at least a year.

Using similar parameter values to those used byFerreira et al. (2002), we reproduce the papillary tumorresults from the same paper to provide a base point for alater consideration of the effects of the immune system.In relation to Fig. 2, the coefficients in the nutrientPDEs have been changed such that the rate ofconsumption by tumor cells of the mitosis nutrient isdoubled, while the consumption rate of the survivalnutrient is decreased by more than half. This allows thetumor cells to divide more rapidly in the direction of thenutrient supply and leads to the ‘‘branchy’’ nature of theresulting tumor shown in Fig. 3(b). Note also that amuch larger domain size was used in Fig. 2 as when the

domain sizes are smaller (as in Fig. 3), the compacttumor grows quickly to completely cover the domainshown and is less rounded in shape. We have useddifferent domain sizes in order to best show the growthpattern of the two tumor types, prior to vascularization.

Fig. 3(a) shows the tumor cell count over time and itcan be observed that the tumor is growing exponentiallythroughout the time considered without moving to alinear growth rate (as is seen in Fig. 2(a)). It appears thatthis is due to the shape of the tumor and the lowerrequirements of the tumor cells for survival nutrient.Unlike the spherical tumors for which the cell-denseperiphery limits the diffusion of nutrients to the tumorcenter, the papillary tumor exhibits a fast expansionfrom its origin and does not form a cell-dense periphery.Nutrient diffusion throughout the domain is easier andmore cells are provided with the nutrients to bothsurvive and divide.

4.2. The effects of the immune system

In this section we investigate the changes to tumorgrowth when an immune system is introduced to themodel. A review of relevant literature suggests that anappropriate range of values for I0, the normal level ofNK cells, is quite broad. For example Kaufmann (inLin, 2004) suggests that the lymphocyte to tumor cellratio can be anywhere from 5:1 to 1:100, depending onthe tumor cell line. Cerwenker and Lanier (2001) statethat up to 15% of lymphocytes are natural killer cells,and with lymphocytes comprising 1012 of the humanbody’s 1013–1014 cells (Encyclopædia Britannica, 2004),this gives a range for I0 of between 0:1% and 1%.

ARTICLE IN PRESS

0 100 200 300 400 500 600 700 8000

1

2

3

4

5

6

7x 105 Tumor cell population over time

Total tumor cell count over time. Tumor cell distribution over the cellular automata grid.

Tumor cell cycles

Tum

or c

ell c

ount

1000

900

800

700

600

500

400

300

200

100

100 200 300 400 500 600 700 800 900 10000

1

2

3

4

5

6Tumor cell distribution after iteration

(a) (b)

Fig. 2. Compact tumor growth in the absence of immune system interaction. Parameter values are: domain size of 1000 elements ! 10220mm,tend " 800 cell division cycles, ynec " 0:03, ydiv " 0:3, ymig " 1000, ln " 50, lm " 25, a " 1, I0 " 0. Note the beginning of a necrotic core in (b).

D.G. Mallet, L.G. De Pillis / Journal of Theoretical Biology 239 (2006) 334–350342

Page 10: A cellular automata model of tumor–immune system interactionszapata/Journalclub/Mallet_2005.pdf · Journal of Theoretical Biology 239 (2006) 334–350 A cellular automata model

Throughout the simulations to be presented we willemploy a value of I0 ! 0:01 or 1%. Values above thiswere found to swamp the early development of tumors,and would therefore be indicative of an immune systemof quite effective strength. For immune levels below 1%,little effect due to the immune system was evident, andwhile this would be indicative of a very weak immunesystem we do not investigate these possibilities due totheir similarities to the immune system-free simulationsof Section 4.1.

Fig. 4 show the effect on tumor (left figures) and totalimmune (right figures) cell populations, that is the sumof NK and CTL cell populations, due to changes in theCTL recruitment parameter yL. All figures wereproduced with the same parameters as Fig. 2 (exceptfor the immune parameters) and show oscillatorypopulation cell counts for both tumor and total immunecells. While these results apply directly to the simula-tions using compact-tumor parameters, qualitativelysimilar results were observed for simulations using thepapillary-tumor parameters.

Figs. 4(a) and (b) show solutions for the lowest valueof yL, when recruitment of CTLs following NK-inducedtumor cell apoptosis is high, and very few oscillationsare observed. The tumor grows exponentially to around15 cells before immune recognition which decreases thetumor size. The tumor begins another growth stage atiteration 25, as it has undergone sufficient shrinkage toevade the immune cells present at that time. The tumorgrows to 35 cells, before the immune system againdetects it and fully eradicates the growth at iteration 42.While it appears in Fig. 4(b) that little has changed inthe immune cell population, the important factor is thelocation of the immune cells. After detection of thetumor, the immune cells are attracted to the location ofthe tumor mass, thus aiding in its removal.

For yL ! 5, solutions are shown in Figs. 4(c) and (d).The tumor cell population is quite oscillatory and isalmost reduced to zero on a number of occasions.However, over time the tumor cell population grows toa somewhat steady level, oscillating between 150 and550 cells. A number of simulations carried out with thisparameter set also exhibited this almost stable, oscilla-tory behavior. These results are similar to the purelytemporal results of Kirschner and Panetta (1998), whereoscillations were also observed in effector and tumor cellpopulations. Experimental evidence for such oscillatorybehavior can also be found in works such as that due toKrikorian et al. (1980) regarding non-Hodgkin’s lym-phoma.

In Figs. 4(e) and (f) the tumor and total immune cellpopulations are shown for the case where CTLinduction is low. In this example, the tumor cellpopulation is only slightly oscillatory. Furthermore,the population grows exponentially (although at aslower rate than in Fig. 2) for the majority of thesimulation. Near the end of the observed period of time,the tumor cell population begins to oscillate around14,000 cells. In other simulations using this parameterset, the low recruitment of T cells leads to the tumorundergoing unstable, oscillatory growth where itappears to be growing without any bounds due to theimmune system. While the stochastic nature of themodel leads to difficulties in determining when thesolution behaviors change, simulations indicated thattumor growth becomes unstable for values of the CTLrecruitment parameter near yL ! 5:5.

Fig. 5 show the effect on tumor (left figures) and totalimmune (right figures) cell populations due to changes inthe CTL death parameter yLD. All figures were producedwith the same parameters as the compact tumor inFig. 2 (except for the immune parameters) and again,

ARTICLE IN PRESS

Fig. 3. An example of papillary tumor growth in the absence of immune system interaction. Parameter values are: domain size of 250 elements" 2:524mm, tend ! 597 cell cycles, ynec ! 0:03, ydiv ! 0:3, ymig ! 1000, ln ! 100, lm ! 10, a ! 2, I0 ! 0.

D.G. Mallet, L.G. De Pillis / Journal of Theoretical Biology 239 (2006) 334–350 343

Page 11: A cellular automata model of tumor–immune system interactionszapata/Journalclub/Mallet_2005.pdf · Journal of Theoretical Biology 239 (2006) 334–350 A cellular automata model

oscillatory population cell counts are observed for bothtumor and the total population of immune cells. Notethat qualitatively similar results were also observed forsimulations using the papillary-tumor parameters,although the change from stable-oscillatory to un-stable-oscillatory growth occurred at a slightly higheryLD value for the papillary tumors.

Figs. 5(a) and (b) show solutions for the lowest valueof yLD, indicating lower probabilities of CTL death inregions of low tumor presence. The solution shows atumor cell population with basically two oscillationsbefore eradication. Note that after the first response bythe immune system, a 50 cell tumor is reduced to onlyone cancerous cell. Exponential growth again occursbefore a complete eradication of the tumor by theimmune cells.

The second row of figures (Figs. 5(c) and (d)) showtumor and total immune cell populations over time foryLD ! 0:5. That is, a higher probability of CTL death

when tumor cell levels are low. Again the tumor cellpopulation oscillates, although about six times longerthan the previous example and with tumor populationsapproximately six times greater in the largest oscilla-tions. This tumor is however, eventually destroyed bythe response of the immune system.

In the final pair of solution plots, Figs. 5(e) and (f),unstable oscillatory tumor growth is evident. Here theCTL death parameter is set to yLD ! 0:7, which is asubstantially lower death probability than for the firstpair of Fig. 5 where tumor destruction was quite fast. Inthis simulation, it appears that initially the tumor is in aphase of somewhat stable, oscillatory growth. However,near the end of the simulation (where the tumor reachedthe edge of the spatial domain), the tumor cellpopulation appears to have broken out of this stablepattern and is growing in an unstable manner. Simula-tions indicated that tumor growth for compact-tumorsimulations becomes unstable when the CTL death

ARTICLE IN PRESS

0 10 20 30 400

10

20

30

40Tumor cell population over time

Tumor cell cycles!L = 3. !L = 3.

!L = 5. !L = 5.

!L = 7. !L = 7.

Tum

or c

ell c

ount

0 10 20 30 40580

600

620

640

660

Immune cell population over time

Tumor cell cycles

Imm

une

cell

coun

t

0 100 200 300 400 5000

100

200

300

400

500

600Tumor cell population over time

Tumor cell cycles

Tum

or c

ell c

ount

0 100 200 300 400 500560580600620640660680700720

Immune cell population over time

Tumor cell cycles

Imm

une

cell

coun

t

0 100 200 300 4000

5000

10000

15000Tumor cell population over time

Tumor cell cycles

Tum

or c

ell c

ount

0 100 200 300 400

1000

2000

3000

4000

5000Immune cell population over time

Tumor cell cycles

Imm

une

cell

coun

t

(a) (b)

(c) (d)

(e) (f)

Fig. 4. The effect on tumor and total (NK+CTL) immune cell distributions in compact tumors over time due to changes in CTL recruitment rates(i.e. changes in yL). Parameter values are: domain size of 250 elements " 2:524mm, ynec ! 0:03, ydiv ! 0:3, ymig ! 1000, ln ! 100, lm ! 10, a ! 2,I0 ! 0:01, yLD ! 0:3.

D.G. Mallet, L.G. De Pillis / Journal of Theoretical Biology 239 (2006) 334–350344

Page 12: A cellular automata model of tumor–immune system interactionszapata/Journalclub/Mallet_2005.pdf · Journal of Theoretical Biology 239 (2006) 334–350 A cellular automata model

parameter was set near yLD ! 0:58, while for papillary-tumor parameters the critical parameter value wasapproximately yLD ! 0:63.

4.3. Lymphocyte infiltration

In this section we present a particularly interestingapplication of the model which produces simulationsthat are qualitatively similar to the results of somerecent experimental studies. Recent experimental studieshave discussed the relationship between increasedsurvival rates of cancer patients, tumor necrosis andfibrosis, and the presence of intratumoral T cells, orinfiltrated T lymphocytes (Schmollinger et al., 2003;Soiffer et al., 1998; Zhang et al., 2003). The resultsshown in Fig. 6 simulate the infiltration of immune cellsinto a growing tumor. These are seen in the darkerregions of Fig. 6(b) where tumor cell necrosis hasoccurred, and in the lighter regions of Fig. 6(d) where

the immune cell numbers are highest. These solutionplots are similar to experimental results shown bySchmollinger et al. (2003), Soiffer et al. (1998), andZhang et al. (2003) where strings of immune cells aremoving into the tumor, surrounding individual cells andcausing tumor cell necrosis.

The simulation shown uses parameters for a compacttumor (in the absence of the immune system), lowlevel CTL recruitment and low CTL death proba-bility. Here we note that the tumor is able to continue togrow in size even though its structure is depletedin a random manner and necrotic regions arevisible throughout the tumor. Essentially, the immunesystem is chasing the tumor but unable to successfullycontain it due to the excessive proliferation of thetumor cells.

A number of experimental studies have producedsimilar patterns of tumor infiltration by immune cells asthose found through solution of the mathematical model

ARTICLE IN PRESS

0 10 20 30 40 500

10203040506070

Tumor cell population over time

Tumor cell cycles

Tum

or c

ell c

ount

0 10 20 30 40 50560580600620640660680700

Immune cell population over time

Tumor cell cycles

Imm

une

cell

coun

t

0 50 100 150 200 250 300 3500

100

200

300

400Tumor cell population over time

Tumor cell cycles

Tum

or c

ell c

ount

0 50 100 150 200 250 300 350

600

650

700

Immune cell population over time

Tumor cell cyclesIm

mun

e ce

ll co

unt

0 100 200 300 400 5000

200

400

600

800

1000

1200Tumor cell population over time

Tumor cell cycles

Tum

or c

ell c

ount

0 100 200 300 400 500550

600

650

700

750

800Immune cell population over time

Tumor cell cycles

Imm

une

cell

coun

t

(a) (b)

(c) (d)

(e) (f)

!LD = 0.3. !LD = 0.3.

!LD = 0.5. !LD = 0.5.

!LD = 0.7. !LD = 0.7.

Fig. 5. The effect on tumor and total (NK+CTL) immune cell distributions in compact tumors over time due to changes in CTL death rates (i.e.changes in yLD). Parameter values are: domain size of 250 elements " 2:524mm, ynec ! 0:03, ydiv ! 0:3, ymig ! 1000, ln ! 50, lm ! 25, a ! 2,I0 ! 0:01, yL ! 3.

D.G. Mallet, L.G. De Pillis / Journal of Theoretical Biology 239 (2006) 334–350 345

Page 13: A cellular automata model of tumor–immune system interactionszapata/Journalclub/Mallet_2005.pdf · Journal of Theoretical Biology 239 (2006) 334–350 A cellular automata model

developed in this paper (see for example, Schmollingeret al., 2003; Soiffer et al., 1998; Zhang et al., 2003).Shown in Fig. 6(e) are patterns of T cell infiltration inovarian carcinoma (taken from Zhang et al. (2003)).Tumor cells (blue) have been extensively infiltrated by

the immune cells (gray). In Fig. 6(f ), a colon metastasis(blue cells) has been extensively infiltrated by CD8+ Tcells (black) (taken from Schmollinger et al. (2003)).Small areas of tumor necrosis (white) can be seen in theregion of the immune cells.

ARTICLE IN PRESS

Fig. 6. Immune cell infiltration into a growing tumor. (a) and (b) show the tumor cell count over time and the final tumor cell distribution over thecellular automata grid, while the same outputs for total immune cells are shown in (c) and (d). Parameter values are: domain size of 250 elements! 2:524mm, tend " 354 cell division cycles, ynec " 0:03, ydiv " 0:3, ymig " 1000, ln " 50, lm " 25, a " 1, I0 " 1%, yL " 9, yLD " 0:9. Computationswere halted when the tumor cells reached the edge of the computational domain. (e) and (f) show T cell infiltration in ovarian carcinoma (from Zhanget al. (2003) used with permission, copyrightrMassachusetts Medical Society, 2003) and colon metastasis (from Schmollinger et al. (2003) used withpermission, copyright r National Academy of Sciences, 2003), respectively.

D.G. Mallet, L.G. De Pillis / Journal of Theoretical Biology 239 (2006) 334–350346

Page 14: A cellular automata model of tumor–immune system interactionszapata/Journalclub/Mallet_2005.pdf · Journal of Theoretical Biology 239 (2006) 334–350 A cellular automata model

Fig. 7 show the infiltration of T cells into the growingtumor over time and the resulting tumor destruction. Inthis simulation, the greater T cell recruitment and lowerdeath rate (compared with the parameters for Fig. 6)cause an early and strong immune response to the youngtumor, as seen in Fig. 7(a)–(e). Over the remainder of

the period of growth, up until total destruction atiteration 256, the tumor cell population undergoesoscillatory growth. The tumor cells grow away fromthe immune cells in an attempt to evade them, while theimmune cells are constantly directed to surround andkill the tumor cells. Eventually the cell division of the

ARTICLE IN PRESS

Fig. 7. Six time frames of immune cell infiltration into a growing tumor leading to tumor destruction. Heavy damage is inflicted early in the tumorgrowth process. Parameter values are: domain size of 250 elements ! 2:524mm, tend " 256 cell division cycles, ynec " 0:03, ydiv " 0:3, ymig " 1000,ln " 50, lm " 25, a " 1, I0 " 1%, yL " 4, yLD " 0:4.

D.G. Mallet, L.G. De Pillis / Journal of Theoretical Biology 239 (2006) 334–350 347

Page 15: A cellular automata model of tumor–immune system interactionszapata/Journalclub/Mallet_2005.pdf · Journal of Theoretical Biology 239 (2006) 334–350 A cellular automata model

tumor cells is not sufficient to exceed the level of celllysis by immune cells and the tumor is destroyed.

5. Conclusion

We have employed a moderately complex, hybridcellular automata-partial differential equation model todescribe interactions between a growing tumor and theimmune system.

The model includes both the growth dynamics oftumors and the effects on tumor growth of the innateand specific immune systems. Each tumor cell is giventhe opportunity to move or divide, and is killed whenthe local survival nutrient levels are too low. Theimmune system is modeled by the inclusion of bothnatural killer cells and cytolytic T lymphocytes, both ofwhich are able to lyse cancer cells. While NK cells cankill only one tumor cell, this initial detection of mutatedcells leads to recruitment of specific immune cells(CTLs) that are able to lyse tumor cells many times.The probability of tumor cell lysis by immune cells isincreased for strong local immune systems. A back-ground level of NK cells is maintained over time toallow for the innate immune system’s surveillance role.After lysing a tumor cell, T lymphocytes continue toroam the domain in search of further target cells,although they may move out of the domain of interest ordie if the local tumor cell density is low.

In the absence of the immune system, the model is able toreproduce tumors of both compact-circular and papillarymorphologies. These tumor shapes have direct dependenceon the relative rates of consumption of the survival andmitosis nutrients by both tumor and host tissue cells. Forcircular tumors, the two-dimensional analogue of thespherical tumor, simulations show the presence of necroticcores and outer bands of proliferating cells. These resultscorrespond qualitatively with the experimental literature(such as Folkman and Hochberg, 1973).

Introducing the immune system to the model leads tovarious results depending on the choice of T lymphocyterecruitment and death parameters. For virtually allimmune system parameter choices, the resulting solu-tions displayed oscillatory behavior for tumor andimmune cell populations in a similar manner to ODEmodels of tumor–immune system interactions. Depend-ing on the strength of T cell recruitment and the T celldeath parameter, the tumor grew with stable or unstableoscillations and in some cases the tumor was destroyedcompletely. Due to the probabilistic nature of themodel, it is difficult to determine exact parameter valuesat which the tumor growth becomes unstable. For theparameter sets shown in Figs. 4 and 5, approximatevalues determined through repeated simulation aregiven in Section 4.2.

The varying parameter value sets correspond tostrong immune systems of healthy individuals, capableof early tumor detection and destruction, and weakerimmune systems of individuals for whom tumors groweasily or are at least able to grow to the point ofmetastasis and hence migrate throughout the body.Furthermore, simulations indicate that (as would beexpected) increasing lymphocyte recruitment and in-creasing the cytolytic ability of lymphocytes leads togreater reductions in tumor size. Such lymphocyterecruitment could be effected through vaccinations withirradiated tumor cells, engineered to secrete chemo-attractants such as those used in the work of Schmol-linger et al. (2003).

Finally, we used the model to demonstrate T cellinfiltration into tumor growths. Experimental literaturehas shown that greater survival rates are observed inpatients with lymphocyte infiltrates in tumors. Using themodel proposed in this paper we have determined theparameter regimes in which infiltration is effective intumor destruction. It is the subject of another study ofthe authors’ to model the effects of injected T cells and Tcell infiltration on pre-grown tumors.

There is clinical evidence that immune cells play animportant role in the control of certain malignancies, afact that is exploited in the development of immu-notherapies for cancers (see, for example Blattman andGreeberg, 2004; Couzin, 2002; Donnelly, 2003; Pardoll,1998; Rosenberg et al., 2004; Wheeler et al., 2004). Ofthe large array of immunotherapy types available,one that might be readily included in this modelstructure is the direct injection of tumor infiltratinglymphocytes (see, for example, Dudley et al., 2002). Infuture studies, the model employed here will be extendedto consider the effects of immunotherapy on theinteractions between the immune system and growingtumors.

In summary, while the cellular automata rulesproposed in this model are not definitive choices, wehave successfully proposed a framework for includingthe immune system in a hybrid CA–PDE model oftumor growth. The resulting solutions are in qualitativeagreement with both the experimental and theoreticalliterature, including both PDE models of tumor growthand ODE models of tumor–immune system interactions.It has been shown that the simple description proposedhere regarding immune and tumor cell interactions, andthe use of hybrid CA–PDE models, have the potential toproduce the behavior observed in some experiments.What needs to follow this work is an exploration of therobustness of the model to different functional forms, aswell as the derivation of actual parameters and accuratefunctional forms from experimentalists. With theintroduction of such elements it will be possible tojudge if this is actually how the observed phenomenaoccur and if this modelling strategy is appropriate.

ARTICLE IN PRESSD.G. Mallet, L.G. De Pillis / Journal of Theoretical Biology 239 (2006) 334–350348

Page 16: A cellular automata model of tumor–immune system interactionszapata/Journalclub/Mallet_2005.pdf · Journal of Theoretical Biology 239 (2006) 334–350 A cellular automata model

Acknowledgements

D.G. Mallet thanks Harvey Mudd College for aPostdoctoral Fellowship during which a proportion ofthis work was completed.

References

Adam, J.A., Bellomo, N., 1997. A Survey of Models for Tumor–immune System Dynamics. Birkhauser, Basel.

Alarcon, T., Byrne, H.M., Maini, P.K., 2003. A cellular automatonmodel for tumour growth in inhomogeneous environment.J. Theor. Biol. 225, 257–274.

Anderson, A.R.A., Chaplain, M.A.J., 1998. Continuous and discretemathematical models of tumor induced angiogenesis. Bull. Math.Biol. 60, 877–900.

Araujo, R.P., McElwain, D.L.S., 2004. A history of the study of solidtumour growth: the contribution of mathematical modelling. Bull.Math. Biol. 66, 1039–1091.

Arciero, J.C., Jackson, T.L., Kirschner, D.E., 2004. A mathematicalmodel of tumor–immune evasion and siRNA treatment. DiscreteContinuous Dynamical Systems—Ser. B 4 (1), 39–58.

Bard Ermentrout, G., Edelstein-Keshet, L., 1993. Cellular automataapproaches to biological modeling. J. Theor. Biol. 160, 97–133.

Bellomo, N., Preziosi, N., 2000. Modelling and mathematical problemsrelated to tumor evolution and its interaction with the immunesystem. Math. Comput. Modelling 32, 413–452.

Bellomo, N., Firmani, B., Guerri, L., 1999. Bifurcation analysis for anonlinear system of integro-differential equations modellingtumor–immune cells competition. Appl. Math. Lett. 12, 39–44.

Blattman, J.N., Greeberg, P.D., 2004. Cancer immunotherapy: atreatment for the masses. Science 205, 200–205.

Burden, T., Ernstberger, J., Renee Fister, K., 2004. Optimal controlapplied to immunotherapy. Discrete Continuous DynamicalSystems—Ser. B 4 (1), 135–146.

Burton, A.C., 1966. Rate of growth of solid tumours as a problem ofdiffusion. Growth 30, 159–176.

Byrne, H.M., 1997. The importance of intercellular adhesion in thedevelopment of carcinomas. IMA J. Math. Appl. Med. Biol.305323, 14.

Byrne, H.M., 1999. Using mathematics to study solid tumour growth.In: Proceedings of the Ninth General Meetings of EuropeanWomen in Mathematics, pp. 81–107.

Byrne, H.M., Chaplain, M.A.J., 1995. Growth of nonnecrotic tumoursin the presence and absence of inhibitors. Math. Biosci. 151181,130.

Byrne, H.M., Chaplain, M.A.J., 1996. Modelling the role of cell–celladhesion in the growth and development of carcinomas. Math.Comput. Modell. 24 (12), 1–17.

Byrne, H.M., Chaplain, M.A.J., 1997. Free boundary value problemsassociated with the growth and development of multicellularspheroids. Eur. J. Appl. Math. 8, 639–658.

Callewaert, D.M., Meyers, P., Hiernaux, J., Radcliff, G., 1988.Kinetics of cellular cytotoxicity mediated by cloned cytotoxic Tlymphocytes. Immunobiology 178, 203–214.

Cerwenker, A., Lanier, L.L., 2001. Natural killer cells, viruses andcancer. Nat. Rev. Immunol. October, 41–48.

Couzin, J., 2002. Select T cells, given space, shrink tumors. Science297, 1973.

Donnelly, J., 2003. Cancer vaccine targets leukemia. Nat. Med. 9 (11),1354–1356.

Dormann, S., Deutsch, A., 2002. Modeling of self-organized avasculartumor growth with a hybrid cellular automaton. In: Silico Biol. 2,0035.

Drasdo, D., Loeffler, M., 2001. Individual-based models on growthand folding in one-layered tissues: intestinal crypts and earlydevelopment. Nonlinear Anal. 47, 245–256.

Dudley, M.E., Wunderlich, J.R., et al., 2002. Cancer regression andautoimmunity in patients after clonal repopulation with antitumorlymphocytes. Science 298 (5594), 850–854.

Encyclopædia Britannica. 2004. Cell. Encyclopædia Britannica Online.[Online] http://search.eb.com/eb/article?tocId=9106125.

Ferreira Jr., S.C., Martins, M.L., Vilela, M.J., 2002. Reaction–diffu-sion model for the growth of avascular tumor. Phys. Rev. E 65,021907.

Ferreira Jr., S.C., Martins, M.L., Vilela, M.J., 2003. Morphologytransitions induced by chemotherapy in carcinomas in situ. Phys.Rev. E 67, 051914.

Folkman, J., Hochberg, M., 1973. Self-regulation of growth in threedimensions. J. Exp. Med. 138, 745–753.

Franks, S.J., Byrne, H.M., King, J.R., Underwood, J.C.E., Lewis,C.E., 2003. Modelling the early growth of ductal carcinoma in situ.J. Math. Biol. 47, 424–452.

Franks, S.J., Byrne, H.M., Mudhar, H., Underwood, J.C.E., Lewis,C.E., 2003. Modelling the growth of comedo ductal carcinoma insitu. Math. Med. Biol. 20, 277–308.

Franks, S.J., Byrne, H.M., Underwood, J.C.E., Lewis, C.E., 2005.Biological inferences from a mathematical model of comedo ductalcarcinoma in situ of the breast. J. Theor. Biol. 232 (4), 523–543.

Galach, M., 2003. Dynamics of the tumor–immune system competi-tion—the effect of time delay. Int. J. Appl. Math. Comput. Sci. 13(3), 395–406.

Galle, J., Loeffler, M., Drasdo, D. Modeling the effect of deregulatedproliferation and apoptosis on the growth dynamics of epithelialcell populations in vitro. Biophys. J. 88 (1), 62–75.

Gatenby, R.A., Gawlinski, E.T., 1996. A reaction–diffusion model ofcancer invasion. Cancer Res. 56, 5745–5753.

Greenspan, H.P., 1972. Models for the growth of a solid tumor bydiffusion. Stud. Appl. Math. 51, 317–338.

Kansal, A.R., Torquato, S., Harsh IV, G.R., Chiocca, E.A.,Deisboeck, T.S., 2000. Simulated brain tumor growth dynamicsusing a three dimensional cellular automaton. J. Theor. Biol. 203,367–382.

Kirschner, D., Panetta, J.C., 1998. Modeling immunotherapy of thetumor–immune interaction. J. Math. Biol. 37, 235–252.

Kleinsmith, L.J., Kerrigan, D., Spangler, S., 2001. National cancerinstitute: Science behind the news—understanding cancer. [Online]http://press2.nci.nih.gov/sciencebehind/cancer/cancer01.htm

Krikorian, J.G., Portlock, C.S., Cooney, D.P., Rosenberg, S.A., 1980.Spontaneous regression of non-Hodgkin’s lymphoma: a report ofnine cases. Cancer 46, 2093–2099.

Kuznetsov, V.A., 1997. Basic models of tumor–immune systeminteractions—identification, analysis and predictions. In: Adam,J.A., Bellomo, N. (Eds.), A Survey of Models for Tumor–immuneSystem Dynamics. Birkhauser, Basel.

Kuznetsov, V., Knott, G., 2001. Modelling tumor regrowth andimmunotherapy. Math. Comput. Modell. 33 (12/13), 1275–1287.

Lin, A., 2004. A model of tumor and lymphocyte interactions. DiscreteContinuous Dynam. Systems—Ser. B 4 (1), 241–266.

Mallet, D.G., 2004. Mathematical modeling of the role of haptotaxisin tumour growth and invasion. Ph.D. Thesis, QueenslandUniversity of Technology, Brisbane, Australia.

The Mathworks, Inc., 2002. MATLAB Release 13. Natick, MA.

Matzavinos, A., Chaplain, M.A.J., 2004. Mathematical modelling ofthe spatio-temporal response of cytotoxic T-lymphocytes to a solidtumour. Math. Med. Biol. 21, 134.

McDougal, S.R., Anderson, A.R.A., Chaplain, M.A.J., Sheraratt,J.A., 2002. Mathematical modelling of flow through vascularnetworks: implication for tumour-induced angiogenesis andchemotherapy strategies. Bull. Math. Biol. 64, 673702.

ARTICLE IN PRESSD.G. Mallet, L.G. De Pillis / Journal of Theoretical Biology 239 (2006) 334–350 349

Page 17: A cellular automata model of tumor–immune system interactionszapata/Journalclub/Mallet_2005.pdf · Journal of Theoretical Biology 239 (2006) 334–350 A cellular automata model

Merrill, S.J., 1982. Foundations of the use of enzyme kinetic analogyin cell-mediated cytotoxicity. Math. Biosci. 62, 219–236.

Owen, M.R., Sherratt, J.A., 1997. Pattern formation and spatiotem-poral irregularity in a model for macrophage–tumour interactions.J. Theor. Biol. 189 (1), 63–80.

Owen, M.R., Sherratt, J.A., 1998. Modelling the macrophage invasionof tumours: effects on growth and composition. IMA J. Math.Appl. Med. Biol. 15, 165–185.

Owen, M.R., Sherratt, J.A., 1999. Mathematical modelling ofmacrophage dynamics in tumours. Math. Models Methods Appl.Sci. 9 (4), 513–539.

Owen, M.R., Byrne, H.M., Lewis, C.E., 2004. Mathematical modellingof the use of macrophages as vehicles for drug delivery to hypoxictumour sites. J. Theor. Biol. 226 (4), 377–391.

Pardoll, D.M., 1998. Cancer vaccines. Nat. Med. Vaccine 4(5(Suppl.)), 525–531.

Patel, A.A., Gawlinski, E.T., Lemiueux, S.K., Gatenby, R.A., 2001. Acellular automaton model of early tumor growth and invasion: theeffects of native tissue vascularity and increased anaerobic tumormetabolism. J. Theor. Biol. 213, 315–331.

Paul, W.E., 2003. Fundamental Immunology, fifth ed. Lippincott,Williams & Wilkins Publishers, Philadelphia, PA.

Perelson, A.S., Mackean, C.A., 1984. Kinetics of cell-mediatedcytotoxicity: stochastic and deterministic multistage models. J.Math. Bio. 170, 161–194.

Perelson, A.S., Weisbuch, G., 1997. Immunology for physicists. Rev.Mod. Phys. 69 (4), 1219–1267.

Pettet, G.J., Please, C.P., Tindall, M.J., McElwain, D.L.S., 2001. Themigration of cells in multicell tumor spheroids. Bull. Math. Biol.63, 231–257.

de Pillis, L., Radunskaya, A., 2001. A mathematical tumor model withimmune resistance and drug therapy: an optimal control approach.J. Theor. Med. 3, 79–100.

de Pillis, L.G., Radunskaya, A., 2003a. A mathematical model ofimmune response to tumor invasion. In: Bathe, K.J. (Ed.),Proceedings of the Second MIT Conference on ComputationalFluid and Solid Mechanics, Computational Fluid and SolidMechanics.

de Pillis, L.G., Radunskaya, A., 2003b. The dynamics of an optimallycontrolled tumor model: a case study. Math. Comput. Modell. 37,1221–1244.

Please, C.P., Pettet, G.J., McElwain, D.L.S., 1999. Avascular tumourdynamics and necrosis. Math. Models Methods Appl. Sci. 9,569–580.

Preziosi, L. (Ed.), 2003. Cancer Modelling and Simulation. Taylor &Francis Ltd., Chapman & Hall/CRC, London, Boca Raton, FL.

Riedel, H., 2004. Models for tumour growth and differentiations. In:Alison, M.R. (Ed.), The Cancer Handbook. Wiley, New York.

Rosenberg, S.A., Yang, J.C., Robbins, P.F., Wunderlich, J.R., Hwu,P., Sherry, R.M., et al., 2003. Cell transfer therapy for cancer:lessons from sequential treatments of a patient with metastaticmelanoma. J. Immunother. 26 (5), 385–393.

Rosenberg, S.A., Yang, J.C., Restifo, N.P., 2004. Cancer immunother-apy: moving beyond current vaccines. Nat. Med. 10 (9),909–915.

Schmollinger, J.C., Vonderhelde, R.H., Hoar, K.M., et al., 2003.Melanoma inhibitor of apoptosis protein (ML-IAP) is a target forimmune-mediated tumor destruction. Proc. Natl Acad. Sci. USA100 (6), 3398–3403.

Seiden, P.E., Celada, F., 1992. A model for simulating cognaterecognition and response in the immune system. J. Theor. Biol. 158(3), 328–357.

Smallbone, K., Gavaghan, D.J., et al., 2005. The role of acidity in solidtumour growth and invasion. J. Theor. Biol. 235 (4), 476–485.

Soiffer, R., Lynch, T., Mihm, M., et al., 1998. Vaccination withirradiated autologous melanoma cells engineered to secrete humangranulocyte macrophage colony-stimulating factor generates po-tent antitumor immunity in patients with metastatic melanoma.Proc. Natl Acad. Sci. USA 95, 13,141–13,146.

Sutherland, R.M., 1988. Cell and environment interactions in tumormicroregions: the multicell spheroid model. Science 240, 177–184.

Takayanagi, T., Ohuchi, A., 2001. A mathematical analysis of theinteractions between immunogenic tumor cells and cytotoxic Tlymphocytes. Microbiol. Immunol. 45 (1), 709–715.

Ward, J.P., King, J.R., 1997. Mathematical modelling of avascular-tumour growth. IMA J. Math. Appl. Med. Biol. 14, 3969.

Ward, J.P., King, J.R., 1999. Mathematical modelling of avascular-tumour growth—II. Modelling growth saturation. IMA J. Math.Appl. Med. Biol. 16, 171211.

Weinberg, R.A., 1996. How cancer arises. Sci. Am. 275 (3), 62.Wheeler, C.J., Asha, D., et al., 2004. Clinical responsiveness of

glioblastoma multiforme to chemotherapy after vaccination. Clin.Cancer Res. 10, 5316–5326.

Xu, Y., 2004. A free boundary problem model of ductal carcinoma insitu. Discrete Continuous Dynam. B 4 (1), 337–348.

Zhang, L., Conejo-Garcia, J.R., Katsaros, D., et al., 2003. Intratu-moral T cells, recurrence, and survival in epithelial ovarian cancer.N. Eng. J. Med. 348 (3), 203–213.

ARTICLE IN PRESSD.G. Mallet, L.G. De Pillis / Journal of Theoretical Biology 239 (2006) 334–350350