Cancer Res-2009-Paiva-1205-11

download Cancer Res-2009-Paiva-1205-11

of 8

Transcript of Cancer Res-2009-Paiva-1205-11

  • 8/2/2019 Cancer Res-2009-Paiva-1205-11

    1/8

    2009;69:1205-1211. Published OnlineFirst January 27, 2009.Cancer ResLeticia R. Paiva, Christopher Binny, Silvio C. Ferreira, Jr., et al.A Multiscale Mathematical Model for Oncolytic Virotherapy

    Updated Version10.1158/0008-5472.CAN-08-2173doi:

    Access the most recent version of this article at:

    MaterialSupplementary

    htmlhttp://cancerres.aacrjournals.org/content/suppl/2009/01/26/0008-5472.CAN-08-2173.DC1.Access the most recent supplemental material at:

    Cited Articleshttp://cancerres.aacrjournals.org/content/69/3/1205.full.html#ref-list-1

    This article cites 30 articles, 11 of which you can access for free at:

    E-mail alerts related to this article or journal.Sign up to receive free email-alerts

    SubscriptionsReprints and

    [email protected] Department atTo order reprints of this article or to subscribe to the journal, contact the AACR

    [email protected] at

    To request permission to re-use all or part of this article, contact the AACR Publications

    American Association for Cancer ResearchCopyright 2009on May 15, 2011cancerres.aacrjournals.orgDownloaded from

    Published OnlineFirst January 27, 2009; DOI:10.1158/0008-5472.CAN-08-2173

    http://cancerres.aacrjournals.org/lookup/doi/10.1158/0008-5472.CAN-08-2173http://cancerres.aacrjournals.org/lookup/doi/10.1158/0008-5472.CAN-08-2173http://cancerres.aacrjournals.org/content/suppl/2009/01/26/0008-5472.CAN-08-2173.DC1.htmlhttp://cancerres.aacrjournals.org/content/suppl/2009/01/26/0008-5472.CAN-08-2173.DC1.htmlhttp://cancerres.aacrjournals.org/content/69/3/1205.full.html#ref-list-1http://cancerres.aacrjournals.org/content/69/3/1205.full.html#ref-list-1http://cancerres.aacrjournals.org/content/69/3/1205.full.html#ref-list-1http://cancerres.aacrjournals.org/cgi/alertshttp://cancerres.aacrjournals.org/cgi/alertsmailto:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]://www.aacr.org/http://www.aacr.org/http://www.aacr.org/http://www.aacr.org/http://www.aacr.org/http://cancerres.aacrjournals.org/mailto:[email protected]:[email protected]://cancerres.aacrjournals.org/cgi/alertshttp://cancerres.aacrjournals.org/content/69/3/1205.full.html#ref-list-1http://cancerres.aacrjournals.org/content/suppl/2009/01/26/0008-5472.CAN-08-2173.DC1.htmlhttp://cancerres.aacrjournals.org/lookup/doi/10.1158/0008-5472.CAN-08-2173
  • 8/2/2019 Cancer Res-2009-Paiva-1205-11

    2/8

    A Multiscale Mathematical Model for Oncolytic Virotherapy

    Leticia R. Paiva,1

    Christopher Binny,2

    Silvio C. Ferreira Jr.,1

    and Marcelo L. Martins1

    1Departamento de Fsica, Universidade Federal de Vicosa, Vicosa, MG, Brazil and 2Centre for Molecular Oncology, Institute of Cancer,Barts and the London School of Medicine and Dentistry, Queen Mary University of London, London, United Kingdom

    Abstract

    One of the most promising strategies to treat cancer is

    attacking it with viruses. Oncolytic viruses can kill tumor cells

    specifically or induce anticancer immune response. A multi-

    scale model for virotherapy of cancer is investigated through

    simulations. It was found that, for intratumoral virus

    administration, a solid tumor can be completely eradicated

    or keep growing after a transient remission. Furthermore,

    the model reveals undamped oscillatory dynamics of tumor

    cells and virus populations, which demands new in vivo and

    in vitro quantitative experiments aiming to detect this

    oscillatory response. The conditions for which each one of

    the different tumor responses dominates, as well as the

    occurrence probabilities for the other nondominant thera-

    peutic outcomes, were determined. From a clinical point of

    view, our findings indicate that a successful, single agent

    virotherapy requires a strong inhibition of the host immune

    response and the use of potent virus species with a high

    intratumoral mobility. Moreover, due to the discrete and

    stochastic nature of cells and their responses, an optimal

    range for viral cytotoxicity is predicted because the virother-

    apy fails if the oncolytic virus demands either a too short or

    a very large time to kill the tumor cell. This result suggests

    that the search for viruses able to destroy tumor cells very

    fast does not necessarily lead to a more effective control of

    tumor growth. [Cancer Res 2009;69(3):120511]

    Introduction

    One of the most promising strategies to treat cancer is attacking

    it with viruses. Oncolytic viruses are biological nanomachines that

    can be programmed to specifically target, replicate in, and

    ultimately kill cancer cells. The new viruses released from the

    lysed cells can subsequently infect adjacent or distant tumor cells

    and trigger multiple cycles of infection. All the recombinant DNA

    technology can be readily used to enhance the selectivity of

    oncolytic viruses for tumor cells, thereby limiting the infection of

    normal cells and minimizing therapeutic side effects (1). Several

    approaches to selectivity have been explored; for example,

    specifically retargeting viruses to cancer cells (2); linking virusreplication to cancer-specific transcription factors such as

    telomerase (3) or prostate-specific antigen (4); and deleting viral

    genes for advancing cell cycle and/or preventing apoptosis,

    typically unnecessary for infecting cancer cells (1). An infected

    tumor cell is eventually killed by the oncolytic virus through severalroutes such as lysis, apoptosis, immune-modulated cell killing, or

    other forms of programmed cell death (5).

    Over the past decade, clinical trials have been conducted with at

    least 14 distinct viruses from six different viral strains (6). Although

    the arsenal of oncolytic viruses expands continuously, to date only

    two viruses (Onyx-015 and H101) have entered randomized phase

    III clinical testing (1, 6). The data from these clinical trials have

    established the safety, effectiveness, and lack of cross-resistance to

    currently available treatments of virotherapy. Hence, in the near

    future, this approach has the potential to supplant many of the

    current standard anticancer therapies, mainly if combined with

    existing treatments. However, several hurdles such as the host

    immune response, systemic distribution, and intratumoral spreadmust be overcome. Because our clinical experience is limited,

    mathematical models may provide relevant insights about the

    nonlinear dynamics ruling the interplay between normal and

    cancer cells and their viral parasites. A quantitative understanding

    of this complex network of interactions will certainly accelerate

    the further development necessary to turn virotherapy into a fully

    functional and effective therapeutics.

    In this article, a virotherapy model recently proposed (7) is

    investigated. It considers normal cells, uninfected and infected

    tumor cells, necrotic cells, and free viruses. At the tissue level,

    reaction-diffusion equations describe the dynamics of nutrients

    and viruses, both small particles in comparison with cells. In these

    reaction-diffusion equations, cells act as sinks and sources ofnutrients and viruses depending on their internal states. In turn,

    cell dynamics is a stochastic process controlled by the local

    concentration of nutrients and free viruses determined at the tissue

    level. Thus, the model involves partial differential equations and

    probabilistic cellular automata rules to describe the multiscale

    dynamics of tumor growth (8, 9). It is worth mentioning that this

    model was designed for the avascular stages of cancer growth.

    However, the effects of therapies even at these stages are crucial for

    the oncologic practice because usually clinically detected tumors

    have often already generated micrometastases that are growing

    avascularly in other tissues of the organism. Thus, for instance, of

    the hundreds of thousand of people who develop colorectal cancer

    every year, as many as 50% of them have synchronous or will havelater hepatic metastases (10). Furthermore, the current available

    treatments offer essentially no hope for cure to those patients with

    unresectable liver metastases.

    Here, the aim is to perform an extensive analysis of the system

    behavior in a subspace of the model parameter space. Our results

    reveal that, for intratumoral virus administration, the solid tumor

    can be completely eradicated or keep growing after a transient

    remission depending on the parameter values. The regions in

    parameter space associated to each one of such tumor responses

    were determined. Finally, our findings are compared with the

    reported results from clinical trials and the predictions of two other

    mathematical models based on systems of ordinary (11) or partial

    Note: Supplementary data for this article are available at Cancer Research Online(http://cancerres.aacrjournals.org/).

    Requests for reprints: Marcelo L. Martins, Departamento de F sica, UniversidadeFederal de Vicosa, 36570-000 Vicosa, MG, Brazil. Phone: 55-31-3899-2993; Fax: 55-31-3899-2453; E-mail: [email protected] American Association for Cancer Research.doi:10.1158/0008-5472.CAN-08-2173

    www.aacrjournals.org 1205 Cancer Res 2009; 69: (3). February 1, 2009

    Research Article

    American Association for Cancer ResearchCopyright 2009on May 15, 2011cancerres.aacrjournals.orgDownloaded from

    Published OnlineFirst January 27, 2009; DOI:10.1158/0008-5472.CAN-08-2173

    http://www.aacr.org/http://www.aacr.org/http://www.aacr.org/http://www.aacr.org/http://www.aacr.org/http://cancerres.aacrjournals.org/
  • 8/2/2019 Cancer Res-2009-Paiva-1205-11

    3/8

    (12) differential equations. Differently from them, the present

    model assumes that each cell is an individual agent and uses a

    sigmoidal cell infection rate (instead of a linear one).

    Materials and Methods

    The mathematical model represents the tissue as a square lattice fed by a

    single capillary vessel at the top of the lattice. As illustrated in Fig. 1, four

    different cell types are considered: normal cells (rn), dead cells (rd), anduninfected (rc) and infected (rv) tumor cells. In contrast to the normal and

    dead cells, one or more uninfected or infected cancer cells can pile up in a

    given site, reflecting the fact that the division of tumor cells is not

    constrained by contact inhibition. Normal cells cannot pile up because they

    are not dividing. Piled, infected cancer cells result from infection of

    uninfected cancer cells in a pile. The nutrients diffuse from the capillary

    vessel through the tissue. Each cancer cell, randomly selected with equal

    probability, can carry out one of three actions: replication, death, or

    becoming infected if yet uninfected. In contrast, an infected cancer cell does

    not divide because its slaved cellular machinery is focused on virus

    replication. It is assumed that infected cancer cells sustain their metabolism

    until lysis. In addition, infected cancer cells mainly die by lysis, and the

    death due to insufficient nutrient supply is neglected. The lysis of each

    infected cancer cell releases v0 viruses to the extracellular medium. Because

    virions are present throughout the infected cell, a random fraction of these

    viruses remains on the site of the lysed cell, and the rest is uniformly

    distributed among its four nearest neighbors at the time of lysis. The

    infection of a normal cell by the oncolytic virus is neglected. Thus, the

    model assumes perfect viral selectivity for cancer cells. At last, the clearance

    rate (cv) embodies the complex innate and adaptive immune responses to a

    virus. Such response involves the synthesis of antiviral cytokines, activation/

    selection of immune cells, and production of antiviral antibodies (13). The

    free viruses diffuse through the tissue. The mathematical details of the rules

    and partial differential equations controlling cellular actions, nutrient

    diffusion, and virus spreading are described in Appendix.

    Figure 1 also illustrates some of the multiple processes involved in the

    model. At the tissue level (macroscopic scale), the nutrient molecules and

    viruses are both very small particles in comparison with cells. Indeed, virus

    species tested in cancer clinical trials form particles with sizes that vary

    from 18 to 30 nm (picornaviruses) to 300 to 400 nm (vaccinia viruses; refs. 1,

    6), whereas a cell is typically 10 Am in diameter. These molecules and

    viruses disperse in the tissue like Brownian particles. Thereby, reaction-

    diffusion equations are used to describe their dynamics. In these equations,

    normal and cancer cells are sinks of nutrients; cancer cells become sinks of

    viruses on infection; and infected tumor cells act as sources of viruses at

    lysis. Thus, the evolution in time of cells determines the macroscopic

    distribution of nutrients and free viruses. In turn, cell division, death,

    infection, and lysis, occurring at the cellular or mesoscopic scale, arestochastic processes controlled by the local concentration of nutrients and

    viruses determined at the macroscopic scale. Such cell actions occur at a

    time scale (f10 h) much greater than that associated to nutrient and virus

    diffusions (f10 s). Hence, cell dynamics evolves in a stationary state of

    these concentration fields.

    Summarizing, both mesoscopic and macroscopic scales are inexorably

    interwoven, although described in terms of distinct physical models: partial

    differential equations (macroscopic level) and probabilistic cellular

    automata rules (mesoscopic level) coupled in a single model through

    division, death, infection, lysis, uptake, release and adsorption rates, and

    diffusion coefficients.

    Results

    We are interested in the effects of virotherapy on solid tumorscommonly used in experimental assays. Figure 2 shows the total

    number of uninfected cancer cells and corresponding virus

    concentrations as functions of time for different virus diffusion

    and clearance rates. The delay between virus and uninfected tumor

    cell dynamics is due to the period for cell lysis (Tl). Depending on

    the parameter values (Table 1), the tumor can be completely

    eradicated few time steps following a single intratumoral virus

    administration, undergo damped oscillations leading to eradication

    at a later time, or sustain an overall tumor growth. In this case, after

    a transient shrinkage of the neoplastic mass, the growth can be

    oscillatory and at a reduced rate or monotonous (nonoscillatory)

    and even at an increased rate. Without treatment, the progress in

    Figure 1. The model considers normal cells, tumor cells, and oncolytic viruses. Uninfected cancer cells can replicate by mitosis or die due to starvation (lack ofnutrients) or be infected by viruses. Such cell actions are stochastic and governed by probabilities Pdiv, Pdel, and Pinf dependent on the nutrient and virus concentrationsper cancer cell. Analogously, normal and infected cancer cells die due to starvation and lysis with probabilities Pdel and Plysis, respectively. The functional formsof such cell action probabilities are stated in Appendix. After mitosis, the new tumor cell piles up (loss of contact inhibition) at the site of its parent cell if the last is insidethe tumor. Otherwise, if the dividing cell is on the tumor border, its daughter cell will randomly replace one of their normal or necrotic nearest neighbor cells. Nutrients andviruses are modeled as continuous concentrations that evolve in space (tissue) and time according to reaction-diffusion equations described in Appendix. Essentially,nutrients diffuse at a rate D from the capillary vessel through the tissue and are taken up by normal and tumor cells at rates c and kc , respectively. In addition,free viruses diffuse at a rate Dv, are cleared at a rate cv, and have as source infected cancer cells, which release v0 new viruses after their lysis. The virus clearancerate represents the simplest way to model viral loss due mainly to the innate and adaptive immune responses mediated by antibodies, CD8 T cells, IFNs, andother cytokines or inherent viral instability. It can also take into account other mechanisms such as the virus outflow from the tumor driven by the gradient ofinterstitial fluid pressure and eventual virus trapping by the extracellular matrix. cv is always a constant, but several distinct values were simulated.

    Cancer Research

    Cancer Res 2009; 69: (3). February 1, 2009 1206 www.aacrjournals.org

    American Association for Cancer ResearchCopyright 2009on May 15, 2011cancerres.aacrjournals.orgDownloaded from

    Published OnlineFirst January 27, 2009; DOI:10.1158/0008-5472.CAN-08-2173

    http://www.aacr.org/http://www.aacr.org/http://www.aacr.org/http://www.aacr.org/http://www.aacr.org/http://cancerres.aacrjournals.org/
  • 8/2/2019 Cancer Res-2009-Paiva-1205-11

    4/8

    time of cancer cell populations for all the simulated patterns is

    Gompertzian (9, 14). The curves in Fig. 2 C and D correspond to

    failed therapies because the cancer keeps growing with time.

    A striking and novel result shown in Fig. 2 is the oscillatorybehavior of both cancer cells and viruses seen at higher viral

    diffusivity and clearance rates. It contrasts to the long-term,

    spatially uniform profiles for cells and free viruses obtained by Wu

    and colleagues (15). Instead of only damped oscillations, the total

    numbers of viruses and uninfected and infected cancer cells can

    also undergo stable aperiodic oscillations superimposed on a

    steady growth.

    In Fig. 3, snapshots of the simulated spatial patterns of

    uninfected tumor cells are shown. The parameter values used are

    those for the curve C in Fig. 2. Successive waves of virus infection

    and therefore cell death by lysis are neatly evident. However, these

    rounds of infection are not strong enough to eradicate the tumor.

    Videos S1 and S2 for unsuccessful and successful treatments,

    respectively, are provided as supplementary information.

    A fundamental feature uncovered by the simulations of the

    model is its great sensitivity to the local, stochastic fluctuations ofuninfected cell population and free virus concentration. Thus, for

    instance, although the complete eradication of the solid tumor is

    the most frequent therapeutic outcome, the other two types of

    tumor response are also observed for cv = 0.01, Dv = 0.7, and Tl = 4.

    Hence, it becomes imperative to find the dominant tumor response

    in every region of the model parameter space. For a given tumor

    cell type (uinf and Tl fixed), the dominant response was determined

    for each pair of parameters Dv and cv controlling free virus

    spreading. Whatsoever the values of Dv and Tl, the therapy always

    fails if the virus clearance rate is large (cvz 0.03). However, a small

    decrease in cv with all the other parameters (Dv, Tl, and u inf) fixed

    can lead to complete tumor eradication (see Supplementary

    Figure 2. Temporal evolution of the number of uninfected cancer cells (Nc) and total free virus concentration with fixed virus infectivity u inf = 0.03 and clearance ratecv = 0.01. A single intratumoral virus load is administered at the time indicated by the arrow. The following virus parameters were used: Dv = 0.7 and Tl = 4, rapideradication (A ); Dv = 0.8 and Tl = 2, slow oscillatory eradication (B); Dv = 0.2 and Tl = 2, oscillatory growth at a reduced rate (C); and Dv = 0.9 and Tl = 32, exponentialgrowth (D). These curves represent the most probable behavior observed for the corresponding sets of model parameters. Each one refers to a single simulation.In all of them, the viruses inside infected cells are not counted for determining free virus concentration in the system.

    Multiscale Model for Virotherapy

    www.aacrjournals.org 1207 Cancer Res 2009; 69: (3). February 1, 2009

    American Association for Cancer ResearchCopyright 2009on May 15, 2011cancerres.aacrjournals.orgDownloaded from

    Published OnlineFirst January 27, 2009; DOI:10.1158/0008-5472.CAN-08-2173

    http://www.aacr.org/http://www.aacr.org/http://www.aacr.org/http://www.aacr.org/http://www.aacr.org/http://cancerres.aacrjournals.org/
  • 8/2/2019 Cancer Res-2009-Paiva-1205-11

    5/8

    Fig. S1). Thereby, virus clearance rate is a key factor determining

    therapeutic outcome. In turn, in a tissue characterized by

    a moderate clearance rate (0.01 V cv < 0.03), ifu inf z 0.1 (i.e., the

    virus has a low efficacy to invade uninfected cancer cells), the

    tumor mass steadily grows and, again, the treatment fails. In

    Fig. 4, the boundaries corresponding to the various therapeutic

    outcomes of an aggressive virotherapy are shown.

    Finally, in addition to determining the most probable prognosis

    after virotherapy, it is relevant to know the chances for the

    occurrence of the other tumor responses in every region in the

    parameter space. This is especially true for the probability of

    therapeutic success. For the aggressive treatment considered here,those probabilities are shown in Fig. 5.

    Discussion

    Computer simulations of our multiscale model revealed that,

    after intratumoral virus administration, a solid tumor can

    completely be eradicated or keep growing after a transient

    remission. Furthermore, depending on the parameter values, the

    tumor cell and virus populations exhibit undamped oscillatory

    dynamics around increasing, nonstationary values (Fig. 2). The

    most likely therapeutic responses are determined by the oncolytic

    activity (u inf and Tl) and spreading properties (Dv and cv) of the

    virus. These are the key parameters that control treatment success.

    Figure 3. Snapshots of cancer cell patterns at intervals of 3 time steps after virus injection. The first pattern corresponds to the solid tumor immediately before thetreatment. The parameters are the same used for the oscillatory growth at a reduced rate shown in Fig. 2 C. Such spatiotemporal evolution can hardly be observedexperimentally either in vitro or in vivo. Indeed, histologic samples require killing of the cells, and noninvasive methods (PET/fluorescently labeled virus) oftenprovide fuzzy pictures characterized by a low spatial accuracy.

    T ab le 1. Parameter values used for a compactmorphology (udiv = 0.3, udel = 0.03 without cell migration)

    Parameter Description Value

    D Glucose diffusion constant 3.6 101 mm2 h1

    c Nutrient uptake rate of

    normal cells

    1065 104 s1

    k cc/c 10Dv Virus diffusion coefficient 10

    2 mm2 h1

    cv Virus clearance rate 2.5 102 h1

    Tl Characteristic time for cell lysis 20 h

    u inf 0.911 MOI, assuming Pinf = 70%

    just after intratumoral injection

    MOI e [0.01,0.1]

    D Linear cell size (lattice constant) 10 Am

    Abbreviation: MOI, multiplicity of infection.

    Cancer Research

    Cancer Res 2009; 69: (3). February 1, 2009 1208 www.aacrjournals.org

    American Association for Cancer ResearchCopyright 2009on May 15, 2011cancerres.aacrjournals.orgDownloaded from

    Published OnlineFirst January 27, 2009; DOI:10.1158/0008-5472.CAN-08-2173

    http://www.aacr.org/http://www.aacr.org/http://www.aacr.org/http://www.aacr.org/http://www.aacr.org/http://cancerres.aacrjournals.org/
  • 8/2/2019 Cancer Res-2009-Paiva-1205-11

    6/8

    For example, independently of the viral diffusivity, infectivity, and

    time for cell lysis, a high clearence rate (cv z 0.03) leads to a

    neoplastic mass and an invaded area that grow monotonously and

    even faster than without treatment. Because the immune response

    directed at free virus and viral antigens expressed on the surface of

    infected cancer cells enhances cv, this result indicates the necessity

    of a strong ablation of the patient immune system. Virotherapy also

    fails when the tumor growth is oscillatory and progresses at a small

    rate. This is the most frequent outcome of varied treatments.

    For instance, the use of an oncolytic virus with a low infectivity

    (u infz 0.1) in a tissue characterized by a moderate virus clearance

    rate (0.01 V cvV 0.03). Here, the failure is caused by the inefficiencyof the viruses at infecting tumor cells, not by the strength of the

    host immune response. The same outcome can be obtained in such

    a tissue for oncolytic viral agents with greater infectivity (0.01 V

    u inf < 0.03) but a small diffusivity (Dv V 0.2; Fig. 3). Now, instead of

    the host immune response or virus inefficacy to infect tumor cells,

    the therapeutic failure is due to infection waves expanding slower

    than the neoplastic mass. Moreover, because a fraction of the

    tumor cells are killed by the virus, the nutrient supply available for

    the uninfected cells becomes larger. In response, they divide more

    frequently, spread, and invade a larger area of the tissue.

    With regard to a successful therapy, our results indicate that the

    key factor is really the virus clearance rate. Indeed, in a host tissue

    with small virus clearance rate (0.005 V cv < 0.01), even mildoncolytic viruses (0.03 V u inf < 0.1) with a rapid virus spreading

    (Dv z 0.3) can eradicate the tumor. In agreement, in vivo

    experiments reveal that the potency of oncolytic viruses is

    markedly reduced when moved to an immunocompetent model

    (16). It is also important to note that, in an immunocompentent

    host, the evolving immune response to the virus will markedly

    change the effective viral clearance rate in a complex way that is

    currently difficult to predict (17).

    The time for lysis cannot be either too short (Tl < 2 time steps)

    o r v er y l on g (Tl > 16 t im e ste ps) . I nd ee d, d ue t o t he

    nonhomogeneous virus dynamics, new infection waves cannot

    reach the tumor border because either they are infrequent and thus

    start far from this border or the large voids generated inside the

    tumor by rapid cell lysis cannot provide further fuel for such waves.

    Hence, from a clinical point of view, our findings indicate that a

    successful, single agent virotherapy requires a strong inhibition of

    the host immune response and the use of potent virus species with

    Figure 4. Boundaries in parameter space associated to different tumorresponses to an aggressive virotherapy (cv = 0.01, low virus clearance rate;u inf = 0.03, high virus infectivity). In regions labeled A , the cancer is eradicated(absolutely successful treatment); in C, the tumor growth progress at a reducedrate and oscillating, whereas in D the growth is monotonic at a rate that can evenbe greater than that of untreated tumors. In both cases the therapy fails. Thisphase diagram was drawn from simulations involving 200 independentsamples (runs) for 50 points in parameter space arranged as a square lattice.Each simulation lasts for 500 time steps or f3 mo, because a time step is

    estimated as f4 h.

    Figure 5. Probability for tumor eradication (A ), monotonic growth (B), andoscillatory growth (C) as a function of the parameters Tl and Dv. Both u inf = 0.03and cv = 0.01 remain fixed. The probabilities were evaluated from samplesincluding 200 independent, 500 time-step runs for each pair (Tl, Dv) considered.

    Multiscale Model for Virotherapy

    www.aacrjournals.org 1209 Cancer Res 2009; 69: (3). February 1, 2009

    American Association for Cancer ResearchCopyright 2009on May 15, 2011cancerres.aacrjournals.orgDownloaded from

    Published OnlineFirst January 27, 2009; DOI:10.1158/0008-5472.CAN-08-2173

    http://www.aacr.org/http://www.aacr.org/http://www.aacr.org/http://www.aacr.org/http://www.aacr.org/http://cancerres.aacrjournals.org/
  • 8/2/2019 Cancer Res-2009-Paiva-1205-11

    7/8

    a high intratumoral mobility. A rapid immune-mediated clearance

    can be avoided through maintenance or incorporation of virus-

    encoded immune response modifier genes or nongenetic

    approaches such as coating the viruses with polymers (1, 6).

    Possibly, viral genes associated with immune evasion could be used

    to reduce the viral clearance rate (18). In contrast, deliberate up-

    regulation of the immune response to target cancer cells is also

    being considered (19). Approaches to increasing oncolytic efficacy

    include arming the viruses with directly lethal (20) or prodrug-converting (21) enzymes. Enhanced intratumoral diffusivity is

    possible if, for instance, the virus is engineered to incorporate a

    gene encoding for a matrix metalloprotease. This enzyme is

    secreted by the infected cancer cell before its lysis and breaks down

    the surrounding extracellular matrix, facilitating the spreading of

    the viruses released after cell death (22).

    The above scenario about the most frequent therapeutic

    outcomes is a simplified one because their corresponding regions

    in the model parameter space are irregular and intertwined (Fig. 4).

    Furthermore, the model is sensitive to the local, stochastic

    fluctuations of uninfected cell population and free virus concen-

    tration. As a result of such sensitivity, the history of cancer

    progression and therapeutic outcome can drastically vary even fora given set of fixed parameters.

    A relevant issue is how our results compare with those obtained

    from clinical trials (1, 6). Even with intratumoral administration,

    wild or engineered adenoviruses, herpes viruses, and reoviruses

    have characteristically induced only transient tumor remissions (6).

    Thus, for instance, single intratumoral injection of hrR3 virus into

    HT29 flank tumors in nude mice has only reduced the rate of

    monotonic growth of this human colon carcinoma (10), a response

    predicted in Fig. 2D. In addition, transient antitumoral responses

    were observed after single intratumoral administration of Reolysin

    (reovirus) and dl1520 (adenovirus ONYX-015) into superficial (23)

    and head and neck (2426) cancers. Because intratumoral spread

    seems to be inefficient with adenovirus (ref. 6; small diffusivity),

    Figs. 4 and 5 can explain why virotherapy using dl1520 fails.Unfortunately, in these phase I and phase II clinical trials with 11%

    and 14% of positive responses, respectively, it was not reported

    if the tumor growth progress either in a monotonic or oscillatory

    way. A neat indication of such oscillatory tumor growth is observed

    in an ovarian cancer xenograft model attacked with MV-CEA

    (an engineered measles virus) administered intratumorally (27). By

    contrast, vaccinia viruses have induced durable complete responses

    in patients after intratumoral or even i.v. injection. Specifically, the

    BP (28) and the engineered JX-594 (29) vaccinia viruses eradicated

    melanomas in 4 (13%) and 2 (29%) patients, respectively. These

    successful therapeutic outcomes, corresponding to Fig. 2A or B,

    have frequencies consistent with eradication probabilities smaller

    than 35% obtained from the model simulations for Dv V 0.6 andTl > 5 h (Fig. 5A). Again, a long-term oscillatory eradication was not

    evidenced in reported clinical trails. Accurate monitoring of tumor

    burden is often difficult due to irregularly shaped and inaccessible

    tumors. This, combined with the contribution of dead cells to the

    tumor mass, would make detection of oscillating responses very

    difficult. New noninvasive imaging techniques such as high-

    sensitivity fluorescence imaging and positron emission tomography

    (PET) scanning (30) could facilitate quantitative monitoring of

    both living tumor mass and viral load in vivo . This approach could

    be used to test the predictions of our model.

    Very recent results obtained by Nagano and colleagues (31)

    reveal the spatial distribution of herpes simplex virus inside

    tumors developed in mice. These authors show that the induction

    of cancer cell apoptosis before the virus administration enhances

    intratumoral virus spreading and antitumor efficacy by generating

    voids and channels inside the tumor. With the main barriers

    against virus spreading disrupted, intratumoral diffusion is

    likely to be closer to the regimen assumed in our model (i.e.,

    viruses can always diffuse through the neoplastic mass). Indeed,

    the spatial patterns of virus distribution shown in Fig. 2A and B

    of Naganos experiment are very similar to the free virussimulated patterns exhibited in our Supplementary videos S1

    and S2.

    Although the present model does not consider the natural

    barriers to virus spreading (fibrillar collagen in the extracellular

    matrix, narrow space between tumor cells, and contrary

    interstitial pressure gradient) or intratumoral heterogeneity in

    the virus diffusivity (as in necrotic areas, for instance), it can

    predict that virotherapy has a great chance to fail if virus

    diffusivity is very low, in complete agreement with the findings

    of Nagano and colleagues (31). Furthermore, our results indicate

    that virus diffusivity is only part of the puzzle. Larger diffusivity is

    desirable, but a low virus clearance rate and an adequate lysis

    time are also necessary to sustain successive waves of infectionrequired for tumor eradication.

    Finally, our results are compared with those provided by two

    other mathematical models. Wein and colleagues (12) modeled the

    dynamics of oncolytic viruses through a deterministic set of partial

    differential equations. They found that, in the case of a uniform

    virus injection, the tumor mass either grows exponentially or is

    eradicated if the waves of infection can or cannot outpace the

    tumor proliferation. Oscillations in time of the virus population

    were explicitly shown. Furthermore, because the wave speed of the

    infection is inversely proportional to Tl (the time for cell lysis),

    these authors suggest that the most efficacious viruses will be

    those that kill an infected tumor cell faster. In contrast, our results

    indicate that there is an optimal range for Tl. Indeed, if the time

    for cell lysis is very short, the waves of infection cannot control theneoplastic growth. A few isolated cancer cells, distant from the

    large voids produced by rapid cell lysis and plenty of free viruses

    looking for targets, provide the opportunity to tumor escape.

    This fundamental and novel prediction seems to emerge from the

    stochastic nature of our model in which the entry of a virus in a

    cell, the lysis of an infected cell, the replication or death of a

    uninfected cell, etc., are all random, individual events. Cell

    discreteness and stochasticity of cellular responses deeply affect

    the strength of the fluctuations, particularly at tumor regions in

    which the number of cancer cells is very low. Surprisingly, the

    deterministic, ordinary differential equation model proposed by

    Wodarz (11) also predicts an optimal viral cytotoxicity. In this

    model, the infection of a cancer cell occurs only through contactwith an infected cell, which, in turn, is killed by the virus at a rate

    a , the viral cytotoxicity. According this model, if there are

    uninfected cancer cells, an increase in the viral cytotoxicity

    increases the tumor mass because infected cells are eliminated

    before the virus had a chance to significantly spread. The

    reciprocal of a should be interpreted as the time for cell lysis Tl.

    However, this model can only exhibit damped oscillations in tumor

    cell populations.

    In summary, we investigate a multiscale model for virotherapy of

    cancer in which the cells are discrete agents. Our predictions are in

    qualitative agreement with results from clinical reports. Through

    extensive simulations, it was found that, for a single intratumoral

    Cancer Research

    Cancer Res 2009; 69: (3). February 1, 2009 1210 www.aacrjournals.org

    American Association for Cancer ResearchCopyright 2009on May 15, 2011cancerres.aacrjournals.orgDownloaded from

    Published OnlineFirst January 27, 2009; DOI:10.1158/0008-5472.CAN-08-2173

    http://www.aacr.org/http://www.aacr.org/http://www.aacr.org/http://www.aacr.org/http://www.aacr.org/http://cancerres.aacrjournals.org/
  • 8/2/2019 Cancer Res-2009-Paiva-1205-11

    8/8

    virus administration, a solid tumor can completely be eradicated or

    keep growing after a transient remission. Furthermore, the model

    reveals undamped oscillatory dynamics of tumor cells and virus

    populations, which demands for new i n vivo and in vitro

    quantitative experiments aiming to detect this oscillatory response.

    The conditions (regions in the model parameter space) for

    domination of each one of the different tumor responses, as well

    as the occurrence probabilities for the other nondominant

    therapeutic outcomes, were determined. From a clinical point ofview, our findi ngs indicate that a successful, single agent

    virotherapy requires a strong inhibition of the host immune

    response and the use of potent virus species with intratumoral high

    mobility. Moreover, due to the discrete and stochastic nature of

    cells and their responses, an optimal range for viral cytotoxicity is

    predicted because the virotherapy fails if the oncolytic virus

    demands either a too short or a very large time for killing the

    tumor cell. This finding suggests that the virus that kills cancer

    cells most rapidly is not necessarily the more effective agent to

    eradicate the tumor. The implications of such a result for the

    design of new replication-competent viruses are clear.

    Disclosure of Potential Conflicts of Interest

    No potential conflicts of interest were disclosed.

    Acknowledgments

    Received 6/10/2008; revised 8/28/2008; accepted 10/3/2008.Grant support: Brazilian Agencies Coordenacao de Aperfeicoamento de Pessoa l

    de N vel Superior, Conselho Nacional de Desenvolvimento Cient fico e Tecnologico ,and Fundacao de Amparo a` Pesquisa do Estado de Minas Gerais.

    The costs of publication of this article were defrayed in part by the payment of pagecharges. This article must therefore be hereby marked advertisement in accordancewith 18 U.S.C. Section 1734 solely to indicate this fact.

    We thank M.L. Munford (Department of Physics, UFV) for preparing Fig. 1.

    References

    1. Parato KA, Senger D, Forsyth PAJ, Bell JC. Recent

    progress in the battle between oncolytic viruses andtumours. Nat Rev Cancer 2005;5:96576.

    2. Sebestyen Z, de Vrij J, Magnusson M, Debets R,Willemsen R. An oncolytic adenovirus redirected with atumor-specific T-cell receptor. Cancer Res 2007;67:1130916.

    3. Kawashima T, Kagawa S, Kobayashi N, et al. Telomer-ase-specific replication-selective virotherapy for humancancer. Clin Cancer Res 2004;10:28592.

    4. Yu D-C, Chen Y, Seng M, Dilley J, Henderson DR. TheAddition of adenovirus type 5 region E3 enables calydonvirus 787 to eliminate distant prostate tumor xenografts.Cancer Res 1999;59:42003.

    5. Baird SK, Aerts JL, Eddaoudi A, Lockley M, LemoineNR, McNeish I A. Oncolytic adenoviral mutants induce anovel mode of programmed cell death in ovarian cancer.Oncogene 2008;27:308190.

    6. Liu T-C, Galanis E, Kim D. Clinical trials results with

    oncolytic virotherapy: a century of promise, a decade ofprogress. Nat Clin Prac Onc 2007;4:10117.7. Ferreira SC, Jr., Martins ML, Vilela MJ. Fighting cancerwith viruses. Phys A 2005;345:591602.

    8. Martins ML, Ferreira SC, Jr., Vilela MJ. Multiscalemodels for the growth of avascular tumors. Phys LifeRev 2007;4:12856.

    9. Ferreira SC, Jr., Martins ML, Vilela MJ. Reaction-diffusion model for the growth of avascular tumors.Phys Rev E 2002;65:021907.

    10. Yoon SS, Carroll NM, Chiocca EA, Tanabe KK. Cancergene therapy using a replication-competent herpessimplex virus type 1 vector. Ann Surg 1998;228:36674.

    11. Wodarz D. Viruses as antitumor weapons: definingconditions for tumor remission. Cancer Res 2001;61:35017.

    12. Wein LM, Wu JT, Kirn DH. Validation and analysis ofa mathematical model of a replication-competentoncolytic virus for cancer treatment: implications for

    virus design and delivery. Cancer Res 2003;63:131724.13. Weinberg R. The biology of cancer. New York:

    Garland Science; 2007.14. Ferreira SC, Jr., Martins ML, Vilela MJ. Morphology

    transitions induced by chemoterapy in carcinomasin situ. Phys Rev E 2003;67:051914.

    15. Wu JT, Kirn DH, Wein LM. Analysis of a three-wayrace between tumor growth, a replication-competentvirus and an immune response. Bull Math Biol 2004;66:60525.

    16. Coffey MC, Strong JE, Forsyth PA, Lee PWK. Reovirustherapy of tumors with activated Ras pathway. Science1998;282:13324.

    17. White CL, Twigger KR, Vidal L, et al. Characteriza-tion of the adaptive and innate immune response tointravenous oncolytic reovirus (Dearing type 3) during aphase I clinical trial. Gene Ther 2008;15:91120.

    18. Wa ng Y, H al ld en G , H il l R , e t a l. E 3 g en e

    manipulations affect oncolytic adenovirus activity inimmunocompetent tumor models. Nat Biotechnol 2003;21:132835.

    19. Hu W, Davis JJ, Zhu H, et al. Redirecting adaptiveimmunity against foreign antigens to tumors for cancertherapy. Cancer Biol Ther 2007;6:17739.

    20. Fernandez M, Porosnicu M, Markovic D, Barber GN.Genetically engineered vesicular stomatitis virus in genetherapy: application for treatment of malignant disease.J Virol 2002;76:895904.

    21. Foloppe J, Kintz J, Futin N, et al. Targeted delivery ofa suicide gene to human colorectal tumors by aconditionally replicating vaccinia virus. Gene Ther2008;15:136171.

    22. Cheng J, Sauthoff H, Huang YQ, et al. Human matrixmetalloproteinase-8 gene delivery increases the onco-

    lytic activity of a replicating adenovirus. Mol Ther 2007;15:198290.

    23. Morris DG, Forsyth P, Roldan G, et al. A phase I

    clinical trial evaluating intralesional Reolysin (reovirus)in histologically confirmed malignancies [abstract92].Proc Am Soc Clin Oncol 2002;21:a24.

    24. Nemunaitis J, Ganly I, Khury F, et al. Selectivereplication and oncolysis in p53 mutant tumors withONYX-015, an E1B-55 kd gene-delected adenovirus, inpatients with advanced head and neck cancer: a phase IItrial. Cancer Res 2000;60:635966.

    25. Ganly I, Kirn D, Eckhardt SG, et al. A phase I study ofOnyx-015, an E1B attenuated adenovirus, administeredintratumorally to patients with recurrent head and neckcancer. Clin Cancer Res 2000;6:798806.

    26. Nemunaitis J, Khury F, Ganly I, et al. Phase II trial ofintratumoral administration of ONYX-015, a replication-selective adenovirus, in patients with refractory headand neck cancer. J Clin Oncol 2001;19:28998.

    27. Peng KW, Hadac EM, Anderson BD, et al. Pharmaco-kinetics of oncolytic measles virotherapy: eventual

    equilibrium between virus and tumor in an ovariancancer xenograft model. Cancer Gene Ther 2006;13:7328.28. Milton GW, Brown MM. The limited role of

    attenuated smallpox virus in the management ofadvanced malignant melanoma. Aust N Z J Surg 1966;35:28690.

    29. Mastrangelo MJ, Maguire HC, Jr., Eisenlohr LC, et al.Intratumoral recombinant GM-CSF-encoding virus asgene therapy in patients with cutaneous melenoma.Cancer Gene Ther 1999;6:40922.

    30. Weissleder R, Ntziachristos V. Shedding light ontolive molecular targets. Nat Med 2003;9:1238.

    31. Nagano S, Perentes JY, Jain RK, Boucher Y. Cancercell death enhances the penetration and efficacy ofoncolytic herpes simplex virus in tumors. Cancer Res2008;68:3795802.

    Multiscale Model for Virotherapy

    www.aacrjournals.org 1211 Cancer Res 2009; 69: (3). February 1, 2009

    on May 15, 2011cancerres.aacrjournals.orgDownloaded from

    Published OnlineFirst January 27, 2009; DOI:10.1158/0008-5472.CAN-08-2173

    http://www.aacr.org/http://www.aacr.org/http://www.aacr.org/http://cancerres.aacrjournals.org/