ArticlePDF Available

Prediction of Drug Response in Breast Cancer Using Integrative Experimental/Computational Modeling

Authors:

Abstract and Figures

Nearly 30% of women with early-stage breast cancer develop recurrent disease attributed to resistance to systemic therapy. Prevailing models of chemotherapy failure describe three resistant phenotypes: cells with alterations in transmembrane drug transport, increased detoxification and repair pathways, and alterations leading to failure of apoptosis. Proliferative activity correlates with tumor sensitivity. Cell-cycle status, controlling proliferation, depends on local concentration of oxygen and nutrients. Although physiologic resistance due to diffusion gradients of these substances and drugs is a recognized phenomenon, it has been difficult to quantify its role with any accuracy that can be exploited clinically. We implement a mathematical model of tumor drug response that hypothesizes specific functional relationships linking tumor growth and regression to the underlying phenotype. The model incorporates the effects of local drug, oxygen, and nutrient concentrations within the three-dimensional tumor volume, and includes the experimentally observed resistant phenotypes of individual cells. We conclude that this integrative method, tightly coupling computational modeling with biological data, enhances the value of knowledge gained from current pharmacokinetic measurements, and, further, that such an approach could predict resistance based on specific tumor properties and thus improve treatment outcome. [Cancer Res 2009;69(10):4484–92] Major Findings By extracting mathematical model parameter values for drug and nutrient delivery from monolayer (one-dimensional) experiments and using the functional relationships to compute drug delivery in MCF-7 spheroid (three-dimensional) experiments, we use the model to quantify the diffusion barrier effect, which alone can result in poor response to chemotherapy both from diminished drug delivery and from lack of nutrients required to maintain proliferative conditions.
Content may be subject to copyright.
Prediction of drug response in breast cancer using integrative
experimental/computational modeling
Hermann B. Frieboes1, Mary E. Edgerton4, John P. Fruehauf9, Felicity R. A. J. Rose10, Lisa
K. Worrall10, Robert A. Gatenby11, Mauro Ferrari2,3,5,7, and Vittorio Cristini1,3,6,8,*
1School of Health Information Sciences, University of Texas Health Science Center, Houston, Texas
2Division of Nanomedicine, University of Texas Health Science Center, Houston, Texas
3Department of Biomedical Engineering, University of Texas Health Science Center, Houston,
Texas
4Department of Anatomic Pathology, The University of Texas M.D. Anderson Cancer Center,
Houston, Texas
5Department of Experimental Therapeutics, The University of Texas M.D. Anderson Cancer Center,
Houston, Texas
6Department of Systems Biology, The University of Texas M.D. Anderson Cancer Center, Houston,
Texas
7Department of Bioengineering, Rice University, Houston, Texas
8Department of Biomedical Engineering, The University of Texas, Austin, Texas
9Division of Hematology/Oncology, Department of Medicine, University of California, Irvine, Medical
Center, Orange, California
10School of Pharmacy, Centre for Biomolecular Sciences, University Park, University of Nottingham,
United Kingdom
11Departments of Radiology and Integrated Mathematical Oncology, Moffitt Cancer Center, Tampa,
Florida
Abstract
Nearly 30% of women with early stage breast cancer develop recurrent disease attributed to resistance
to systemic therapy. Prevailing models of chemotherapy failure describe three resistant phenotypes:
cells with alterations in transmembrane drug transport, increased detoxification and repair pathways,
and alterations leading to failure of apoptosis. Proliferative activity correlates with tumor sensitivity.
Cell cycle status, controlling proliferation, depends upon local concentration of oxygen and nutrients.
Although physiological resistance due to diffusion gradients of these substances and drug is a
recognized phenomenon, it has been difficult to quantify its role with any accuracy that can be
exploited clinically. We implement a mathematical model of tumor drug response that hypothesizes
specific functional relationships linking tumor growth and regression to the underlying phenotype.
The model incorporates the effects of local drug, oxygen and nutrient concentrations within the three-
dimensional tumor volume, and includes the experimentally observed individual cells’ resistant
phenotypes. By extracting mathematical model parameter values for drug and nutrient delivery from
monolayer (one-dimensional) experiments and using the functional relationships to compute drug
*Correspondence: Vittorio Cristini, University of Texas HSC—SHIS, 7000 Fannin #600, Houston, TX 77030, USA. Tel.: 713-500-3965;
Fax: 713-500-3929; E-mail: E-mail: vittorio.cristini@uth.tmc.edu.
NIH Public Access
Author Manuscript
Cancer Res. Author manuscript; available in PMC 2010 May 15.
Published in final edited form as:
Cancer Res. 2009 May 15; 69(10): 4484–4492. doi:10.1158/0008-5472.CAN-08-3740.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
delivery in MCF-7 spheroid (three-dimensional) experiments, we use the model to quantify the
diffusion barrier effect, which alone can result in poor response to chemotherapy both from
diminished drug delivery and from lack of nutrients required to maintain proliferative conditions.
We conclude that this integrative methodology tightly coupling computational modeling with
biological data enhances the value of knowledge gained from current pharmacokinetic
measurements, and, further, that such an approach could predict resistance based on specific tumor
properties and thus improve treatment outcome.
Keywords
breast cancer; drug response; mathematical model
INTRODUCTION
We implement a novel quantitative approach that links growth and regression of a tumor mass
to the underlying phenotype to study the impact of drug and nutrient delivery as mediators of
physiological resistance. It is well known that inefficient vascularization may prevent optimal
transport of oxygen, nutrients, and therapeutics to cancer cells in solid tumors [1]. As a result,
the drug agent must diffuse through the tumor volume to reach the entire tumor cell population,
and there is mounting evidence to indicate that these diffusion gradients may significantly limit
drug access [2–5]. Both hypoxia and hypoglycemia contribute to physiological resistance
through various mechanisms, including induction of oxidative stress and a decrease in the
number of proliferating cells [6–12]. The myriad of stresses can lead to selection of cells that
resist apoptotic conditions, thus adding to pathologic resistance in tumors [13]. This evidence
strongly suggests that the diffusion process alone can lead to the evolution of drug resistance
in tumor cells that exceeds predictions based on individual cell phenotype [5]. It has not been
easy to quantify the resistance effects of diffusion gradients with any accuracy that can be
exploited in a clinical setting. The different physical scales in a tumor spanning from the nano-
to the centimeter scale present a complex system that to be better understood could benefit
from appropriate mathematical models and computer simulations in addition to laboratory and
clinical observations.
In particular, biocomputational modeling of tumor drug response has endeavored in the last
two decades to address this need. Space limitations preclude a full description (see [5,14,15],
references therein). Dox cellular pharmacodynamics has been modeled; e.g., [16] presented a
model providing good fits to in-vitro cytotoxicity data. Drug transport was modeled in
spheroids vs. monolayers [17]. A model capable of predicting intracellular Dox accumulation
that matched experimental observations was described in [18]. Different drug kinetics effects
in-vitro were compared in [19], showing that a single drug infusion could be more effective
than repeated short applications. Models employing multi-scale approaches, i.e., linking events
at sub-cellular, cellular, and tumor scales (e.g., [20]), studying vascularized tumor treatment
(e.g., [21]), and simulating nanoparticle effects (e.g., [3]) have also been developed. Existing
mathematical models are often limited to radially-symmetric tumor representations and not
fully constrained through experimentally-set parameters.
Here, we employ a multi-scale computational model, extending a previous formulation of
tumor growth founded in cancer biology [22–25], to enable more rigorous quantification of
diffusion effects on tumor drug response. This model can represent non-symmetric solid tumor
morphologies in 3-D, thus providing the capability to capture the physical complexity and
heterogeneity of the cancer microenvironment. More significantly, we fully constrain the
model through functional relationships with parameters set from experiments. We hypothesize
the simplest relationships that would at the same time be biologically founded and which could
Frieboes et al. Page 2
Cancer Res. Author manuscript; available in PMC 2010 May 15.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
be calibrated by the experimental data. These relationships link tumor mass growth and
regression to the underlying phenotype. They provide the mathematical basis for describing
cell mitosis, apoptosis and necrosis modulated by diffusion gradients of oxygen, nutrients, and
cytotoxic drugs, and enable quantification of the physiological resistance introduced by these
gradients. Input parameters include the diffusion coefficients for these substances, and the rate
constants for proliferation, apoptosis, and necrosis. We measure parameter values from
independent experiments performed under conditions with no gradients, i.e., with cells grown
as 1-D monolayers, and then use these values to calculate cell survival in 3-D tumor geometry.
These values are compared with experiments in which cells are grown as in-vitro tumor
spheroids, representing a 3-D tumor environment with diffusion gradients. This approach
allows us 1) to fully constrain the computational model, using experimentally-obtained
parameter values, and 2) to validate the hypothesized functional relationships by comparing
the computed 3-D tumor viability with the spheroid tumor growth experiments. By quantifying
the link between tumor growth and regression and the underlying phenotype, the work
presented here provides a quantitative tool to study tumor drug response and treatment.
Although in-vitro spheroids are a gross simplification of the complex in-vivo condition, they
allow capturing the effect of 3-D tissue architecture in generating diffusion gradients of drug
and cell substrates under controlled conditions without the complicating effects of blood flow
[1]. Spheroids develop a layer of viable cells surrounding a necrotic core [26]. The thickness
of this layer (ca. 100 µm), maintained by substrate diffusion gradients, “mimics” the viable
tissue thickness supported by diffusion from surrounding blood vessels in-vivo [26]. This
provides a more controllable experimental model than in-vivo, which is necessary in order to
test and calibrate the computational model parameters. In turn, computer simulations of the
model enable formulating and testing the functional relationships that quantify the dependence
of drug resistance on diffusion gradients and particular tissue characteristics (e.g., tissue
compactness as a function of cell packing density [27]). We conclude that computational
modeling tightly integrated with tumor biology extends the information that can be learned
from pharmacokinetic experiments [28–30], offering a promising possibility of ultimately
quantifying and predicting treatment response from individual tumor characteristics.
MATERIALS AND METHODS
Doxorubicin was used as the cytotoxic drug in the experiments. Experiments used MCF-7 WT
breast cancer cell lines to represent drug-sensitive tumor cells and MCF-7 40F breast cancer
cell lines to represent drug-resistant tumor cells. Details of experimental methods are given
below.
Observation of cell substrate gradients
Confluent MCF-7 WT (breast cancer) cell monolayers were incubated with 0.25% (w/v) trypsin
(Sigma, UK) and 0.02% (w/v) EDTA (Sigma, UK) for 4 minutes at 37°C. Complete media was
added to this mixture to inhibit enzyme activity and the cell suspension passed through a 24-
gauge needle 6 times to ensure a single cell population. A 10 mL High Aspect Ratio Vessel
(HARV; Synthecon, USA) was seeded at 2×106 cells/mL and rotated at 15 rpm at 37°C in a
humidified atmosphere of 5% CO2 in air. Aggregates formed within 6 hours and were
maintained in RCCS culture for a total of 30 days. During this time, media in the HARVs was
replenished 50:50 every other day. Spheroids were fixed in formalin, embedded in paraffin
and 5 µm sections cut for immunohistochemistry. To observe hypoxia, unfixed spheroids were
incubated in 200 µM pimonidazole (Chemicon, UK) in complete media for 2 hours at 37°C
prior to fixation as above. Proteinase K digestion was used for antigen retrieval prior to 1° Ab
incubation. The mouse IgG 1° Ab was used at 1:600 dilution and incubated with 4 µm sections
for immunolocalization of hypoxia through detection of pimonidazole protein adducts.
Frieboes et al. Page 3
Cancer Res. Author manuscript; available in PMC 2010 May 15.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Immunohistochemical (IHC) staining for GLUT-1 was performed using rabbit polyclonal
antibody against the C-terminal portion (Abcam). The NHE-1 antibody is a rabbit polyclonal
antibody (Santa Cruz). Immunohistochemistry was performed using the Discovery XT
Automated Staining platform (Ventana Medical Systems, Inc, Tucson, Arizona).
Deparaffinization and antigen retrieval of tissue sections was performed online. Antigen
retrieval was performed on the Ventana instrument with a borate buffer (pH 8) 40 minutes at
95°C. Once the tissue was conditioned in this way, primary antibody was applied manually at
a 1:800 dilution. (60 min. incubation at 37°C). Primary antibodies were visualized using VMSI
validated detection and counterstaining reagents. Images were captured using an Olympus
BX50 camera with an RT SPOT (Diagnostic Instruments, Inc), and standardized for light
intensity.
Drug response
MCF-7 WT (drug-sensitive) and MCF-7 40F (drug-resistant) cell lines were cultured in RPMI
Media 1640 (Life Technologies Invitrogen, Carlsbad, California) supplemented with 3% FBS
(Life Technologies Invitrogen, Carlsbad, California), 2 mM L-glutamine, and 1% penicillin/
streptomycin in humidified 7.5% CO2 at 37°C. Cells for monolayer culture were seeded 20,000
per well in 24-well Costar 3527 cell culture plates (Corning, New York). Cells for spheroids
were seeded 50,000 per well in 24-well Costar 3473 cytophobic plates, and shaken at 100 rpm
for 10 min. on day 1. Both the cells in monolayer and spheroids were incubated for 3 days, and
then exposed to doxorubicin (Dox) (Bristol-Myers Squibb, Princeton, New Jersey)
concentrations ranging from 0 to 16384 nM in 4× nM increments (0, 4, 16, 64, etc.) for 96
hours, representing at 256 nM a typical area-under-the-curve1 in-vivo [29]. Negative controls
were seeded and incubated under the same conditions without drug. Three endpoints were
concurrently measured as fraction of negative control: proliferation using tritiated thymidine
(Amersham, Buckinghamshire, Great Britain) incorporation assay [31], viability using trypan
blue exclusion counts, and metabolic activity using XTT (sodium 3'-[1-[(phenylamino)-
carbonyl]-3,4-tetrazolium]-bis(4-methoxy-6-nitro)benzene-sulfonic acid hydrate) assay [32].
All experiments were done in triplicate. Photographs were taken with a digital camera through
a Zeiss microscope (100× magnification).
Mathematical model of tumor growth and drug response
Briefly, the equations [22,23] are formulations of mathematical models used in engineering to
describe phase separation of two partially miscible components (viable and necrotic tumor
tissue), diffusion of small molecules (cell substrates and drug), and conservation of mass
(Quick-Guide, Supplementary Figure 1, see also [33–36]). Mass conservation equations
describe growth (proliferation as a function of total number of cycling cells) and death from
the drug’s cytotoxic effects (apoptosis as a function of a rate constant dependent upon the
unique cell sensitivity and as a function of cycling cells). These are combined with diffusion
of small molecules to a reaction-diffusion equation. Rate constants for proliferation and
apoptosis are modified by functions that represent their dependence upon cell nutrients and
oxygen (proliferation) and drug concentration (death), along with a dependence on spatial
diffusion of these substances. By dimensional analysis of the reaction-diffusion equation the
diffusion penetration length L = (D/η)1/2, where D is the corresponding diffusion constant and
η is a characteristic cellular uptake rate. Large differences in concentration will occur if L is
much less than the tumor radius (linear dimension) in the absence of blood vessels, creating a
steep gradient. By using L and knowing the spheroid size, one can determine the steepness of
these gradients by solving the reaction-diffusion equation. We note that gradients are affected
1In pharmacokinetics, the area-under-the-curve (AUC) is a calculation to evaluate the body's total exposure to a given drug. In a graph
plotting plasma drug concentration vs. time, the area under this curve is the AUC
Frieboes et al. Page 4
Cancer Res. Author manuscript; available in PMC 2010 May 15.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
by cell packing density, intra-cellular uptake, and duration of drug exposure, as well as
conditions specific to particular experiments.
Estimation of mathematical model parameters
In our model, cell substrates are represented by glucose and oxygen with L estimated from
independent measurements taken from the literature. For glucose D ~ 1×107 cm2/sec [37] and
uptake rate may be as large as 1×103 sec1 [38], while for oxygen the corresponding values
are ~ 1×105 cm2/sec [39] and 1×101 sec1 [40]. Diffusion penetration lengths thus calculated
(L 100 µm) are consistent with our experimental observations (IHC) and with previous
(murine mammary tumor EMT6/Ro spheroid) observations showing glucose concentration
decreasing by 65% [41] and oxygen decreasing by 90% [42] across the tumor viable region.
Although gradients of drug into tumor tissue have been observed with a number of different
cell types, many of these observations have been qualitative. Thus, published values for
diffusion constants and uptake rates can vary considerably. We estimate Dox diffusion length
by considering that the distance at which the penetrating concentration will be 50% of the
source is ln 2 times the diffusion length – equivalent L here is ca. 90 microns for a 2/3 drop
across the spheroid viable region (~ 100 µm). This is consistent with in-vitro Dox penetration
into Chinese hamster lung cell spheroids at 24 hours [43] reporting an average 2/3 drop of
external concentration across the viable region. Other studies have shown Dox penetration in-
vivo (mouse model) decreasing by half within 40–50 µm of blood vessels by 3 hours [4], as
well as penetration in humans (biopsies) decreasing by half within 50 µm of blood vessels after
2 hours of a bolus and within 60–80 µm after 24 hours in breast cancer tissue [2]. By fitting
the solution of the unsteady Eq. (3) to this data we could indirectly estimate average cellular
uptake rates η of Dox by breast cancer cells in 3-D tissue of ~ 1 day1 , consistent with the
independent penetration length estimate L = (D/η)1/2 from the steady-state profiles.
The rate λM (inverse time) measures the change in cell number in a population due to mitosis,
normalized by total cell number (control = cells with no drug), and thus was calibrated by
matching proliferation data from our experiments, using as initial guess for λM/λM,C, the in-
vitro cell proliferation as fraction of control divided by in-vitro cell viability (total number of
cells N in monolayer) as fraction of control, yielding an estimate of cell proliferation ~1
day1.
Apoptosis rate λA (inverse time) over the period 0 < t < T (total experiment time) was estimated
as a function of local concentrations of substrate σ and drug d as
and . Death rate λA’ was calculated from measurements of cell
viability N as fraction of negative control NC over time in monolayer cell cultures at various
drug concentrations. Physiological resistance based on cell cycle status was modelled with
λA". Decreased substrate concentration in 3-D results in cell populations that cycle less and
therefore have reduced sensitivity to Dox2. In the above formula, enhanced survival was
assumed linearly dependent on substrate concentration with a fitting constant calibrated from
previous observations [12] showing an average MCF-7 viability increase 2.2 to 4 times for
glucose deprivation of up to 50% at drug concentrations similar to our study, i.e., 2.21+α/
2Cytotoxicity of MCF-7 cells to Dox may be affected more by glucose deprivation than by hypoxia [6,44]. The glucose-regulated stress
response [10] has been correlated with resistance to topoisomerase II-directed chemotherapeutic drugs such as Dox [10,11] through a
decreased expression of topoisomerase II in MCF-7 cells [12].
Frieboes et al. Page 5
Cancer Res. Author manuscript; available in PMC 2010 May 15.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
24.0. In our experiments, glucose concentration in the medium was σ = 2.0 g/L. Note that
for σ = σ, the above formula gives and thus .
Necrosis parameters for diffusion-limited growth λN (rate of necrosis) and σN (substrate limit
for cell viability) were calibrated by matching simulation results to in-vitro spheroid growth
curves and histological data under conditions with no drug, as described previously [45], i.e.,
so that the simulated spheroid radius and viable region thickness matched in-vitro observations.
These parameters are directly responsible for the steady spheroid size (and extent of necrosis)
after a growth period because they regulate the balance of mass growth due to cell proliferation
in the viable region and mass loss due to cell disintegration in the necrotic center [45]. Based
on our in-vitro measurements of spheroid and necrotic volumes we found a stationary average
spheroid radius of ca. 0.8 mm, with viable region of thickness ~0.1 mm (see also [46]). The
latter is consistent with the estimated cell substrate penetration length L calculated above. The
model calibration based on these quantities consistently yielded:λN/λM = 0.7 and σN/σ = 0.5.
Higher cell densities were observed for resistant spheroids, which had a more compact
morphology [27]. Based on these observations, higher cell adhesion parameter values [22]
were used in the model equations to simulate resistant spheroids than for sensitive ones. We
used a recently presented calibration procedure [45] to determine the relative values for
sensitive and resistant cell cultures. This observation and the resulting different parameters for
the two cell phenotypes underscore the need to explicitly model cell adhesion forces when
constructing models of cancer growth.
QUICK-GUIDE TO EQUATIONS AND ASSUMPTIONS
Equation (1)
The 3D in-vitro environment is modeled as a mixture of viable tumor tissue (volume fraction
ϕV), dead tumor tissue (ϕD), and interstitial fluid (ϕW, given indirectly by 1(ϕV+ϕD) and
assumed to move freely) flowing through the extracellular matrix, which we treat as a porous
medium. These partial differential equations are derived from the conservation of mass of these
quantities [22,23]. From left to right in each equation, the terms represent: change of volume
fraction with respect to time; tissue advection (bulk transport) by the overall mixture as it moves
with local velocity u; net tissue diffusion due to the balance of mechanical forces, including
cell-cell adhesion (related to mobility constant M [22,23]), and cell-cell adhesion and repulsion
(oncotic pressure), modeled using the variational derivative of the adhesion energy E (for
specific forms of this energy, see [22]); and net source of tissue due to the balance of mitosis,
apoptosis, and necrosis.
In Words
The temporal rate of change in viable and dead tumor tissue at any location within the tumor
equals the amount of mass that is pushed, transported, and pulled due to cell motion, adhesion,
and tissue pressure, plus the net result of production and destruction of mass due to cell
proliferation and death.
Major Assumptions of the Model
The tumor is a mixture of cells, interstitial fluid, and extracellular matrix (ECM), and we treat
the motion of interstitial fluid and cells through the ECM as fluid flow in a porous medium.
Therefore, no distinction between interstitial fluid hydrostatic pressure and mechanical
Frieboes et al. Page 6
Cancer Res. Author manuscript; available in PMC 2010 May 15.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
pressure due to cell-cell interactions is made. Cell velocity is a function of cell mobility and
tissue oncotic (solid) pressure (Darcy’s law); cell-cell adhesion is modeled using an energy
approach from continuum thermodynamics [22].
Equation (2)
These equations specify net sources SV and SD of mass for viable and dead tumor tissue,
respectively. This directly links the conservation of mass equations Eq. (1) to the cell phenotype
through hypothesized phenomenological functional relationships that include diffusion of cell
substrates and drug (of local concentration σ̣ and d, respectively) through tumor interstitium
(see [22] and references therein). For the viable tissue (first equation), the terms on the right
hand side represent, from left to right, volume fraction gained from cell mitosis (rate λM), and
lost to cell apoptosis (λA) and necrosis (λN). For the dead tissue (second equation), the terms
on the right hand side represent, from left to right, increase in volume fraction from cell
apoptosis and necrosis, and decrease in volume fraction from cell disintegration by lysis (λL).
All rates are inverse time.
In Words
Mass of viable tumor tissue increases through cell proliferation and decreases through cell
apoptosis and necrosis. Mass of dead tumor tissue increases through cell apoptosis and necrosis,
and decreases through cell lysing. These equations provide a means of incorporating the
biology of the problem into the physical conservation laws in Eq. (1).
Major Assumptions of the Model
Cells are composed entirely of water, which is a reasonable first approximation in terms of
volume fraction (see discussion in [22]). Cell mitosis is proportional to substrates present
[47]. As mitosis occurs, an appropriate amount of water from the interstitial fluid is converted
into cell mass. Substrate depletion below a level σN leads to necrosis [26,7]. Cell lysis
represents a loss of mass as cellular membranes are degraded and the mass is converted into
water that is absorbed into the interstitial fluid. The simulated interface between viable and
necrotic tissue is forced to be of biologically realistic thickness (10–100 µm) through the
Heaviside function (see [22,23]. Mitosis and apoptosis rates are functions of drug (d) and
substrate (σ) concentrations. Note that the assumption that parameters measured for drug
sensitivity in monolayer, modulated by diffusion gradients, can adequately represent response
in-vivo may not be correct for real tumors; e.g., it disregards drug resistance from growing with
an ECM.
Equation (3)
This is a partial differential equation that describes cell substrate concentration σ across the
tumor viable region. The first term on the right hand side models diffusion of the substrates
into the tumor spheroid with penetration length L. The second term models substrate uptake
by the cells (with rate η). An analogous equation is posed for drug concentration d.
In Words
The temporal rate of change of cell substrates or drug across the tumor viable region equals
the net amount that diffuses into the region minus the amount uptaken by the tumor cells.
Frieboes et al. Page 7
Cancer Res. Author manuscript; available in PMC 2010 May 15.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Major assumptions of the Model
Diffusion of cell substrates and drug into the tumor, combined with uptake in the tumor interior,
creates and maintains a gradient of these substances through the 3-D tumor tissue, which is
assumed to be fairly compact (non-invasive/infiltrative).
RESULTS
The hypothesized functional relationships of the computational model link growth and
regression of the tumor mass to the underlying phenotype as described in Eq.1–Eq.3 (Quick-
Guide). The model is fully constrained (Methods) through experimentally-based parameters
(Figure 1). In order to quantify physiological resistance in 3-D tumor tissue, diffusion gradients
are incorporated in the functional relationships and the model is calibrated to compute these
gradients. Accordingly, immunohistochemical measurements were performed on drug-
sensitive tumor spheroids cultured in-vitro (Figure 2) to confirm that in our experiments 1) cell
substrate concentration decreases across the radius of the spheroid, 2) apoptosis correlates with
hypoxia, and 3) proliferation correlates with substrate availability [43,47]. Increasing positivity
for pimonidazole protein adducts across the viable rim (~ 100 µm) in the direction of the
spheroid center indicates a gradient of oxygen. Across the viable region, upregulation of Na+/
H+ transporters detected by increasing positivity for NHE-1 is observed, demonstrating a
gradient of pH. Na+/H+ transporters are upregulated in response to acidosis. Upregulation of
GLUT-1 transporters (adaptation to hypoxic/hypoglycaemic stress) is shown in the increasing
positivity for GLUT-1 [48] in peri-necrotic regions, indicating gradients of glucose and oxygen.
The frequency of apoptotic cells increases towards the center as shown by increasing positivity
of TUNEL3 staining. End stage apoptotic bodies are concentrated at the peri-necrotic necrotic
regions with most cells in the center being apoptotic or necrotic. Cell proliferation is limited
to the viable rim, as demonstrated by Ki-67 nuclear staining.
According to the computational model, the solution of the reaction-diffusion Eq. (3) yields
substrate gradients (Figure 3A) developing across a viable region of width ~ 0.1 mm. The
computed width of this region is consistent with previous observations by other investigators
(e.g., [26]). Both glucose and oxygen substrate concentrations were calculated in the model to
drop by at least 50% across the spheroid viable region, with an average concentration σ/σ
0.7 in the living tumor tissue. This is in agreement with measurements of glucose [41] and
oxygen (e.g., [42]) in spheroid viable regions from previously reported in-vitro measurements
(Methods).
Similarly, the model computes the drug gradients that develop within the viable tumor spheroid
tissue. The calculated drug concentration profile reaches steady state in ca. 24 hours and decays
by 2/3 across the viable region to yield an average drug concentration of 50% that of outside
the spheroid (Figure 3B). Note that the drug penetration length is comparable to the penetration
length of substrates (Figure 2). Since Doxorubicin is a cell–cycle specific drug and the
proportion of cells that are cycling correlates with the substrate availability, the drug would be
less effective as the proportion of cycling cells is diminished. Not only is its concentration
decreased by the diffusion barrier, but its efficacy is impaired as a result of the substrate
gradients. The physiologic resistance predicted by the model based on the hypothesized
functional relationships would in principle apply to any cell-cycle specific drug.
Cell proliferation (ratio of proliferation-to-viability counts) in-vitro under conditions with
diffusion gradients (i.e., in spheroids) vs. monolayers was lower by at least 50% for both cell
lines at drug concentrations d > 64 nM representing clinically relevant dosages4 (Figure 4A).
3Terminal deoxynucleotidyl Transferase Biotin-dUTP Nick End Labeling.
Frieboes et al. Page 8
Cancer Res. Author manuscript; available in PMC 2010 May 15.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Further, increased viability for drug-sensitive cells in 3-D corresponded with ca. 50% lower
cell metabolic activity (ratio of metabolic-to-viability counts) in-vitro (data not shown). These
data indicate that cell quiescence was significant at drug dosages similar to in-vivo conditions.
The cell apoptosis parameter of the computational model calculated from the monolayer cell
viability data increases with drug concentration (Figure 4B).
The functional relationships of the computational model (Eq. 1–Eq.3) linking growth and
regression of the tumor mass to the underlying phenotype are validated by quantifying the
physiological resistance introduced by the diffusion gradients. Cell viabilities predicted by the
model agree well with the ones observed experimentally in 3-D spheroids (Figure 5). In
contrast, monolayer cell viability data is a poor prognosticator of drug resistance in 3-D tumors.
An average survival increase (as fraction of control) of 250% over monolayer was predicted
for drug-sensitive (MCF-7 WT) spheroids under various drug concentrations (Figure 5A),
while for drug-resistant (MCF-7 40F) spheroids the corresponding increase was 280% (Figure
5B). The median dose5 for MCF-7 40F was ~20% higher in spheroid vs. monolayer than for
MCF-WT, implying that not only did the 3-D morphology promote the net survival for both
cell types, but also that the resistant phenotype was further favored by the 3-D configuration.
We noted that drug-resistant spheroids maintained more compact, nearly spherical shapes
(Figure 6A) that simply decreased in size as cells were killed, while drug-sensitive spheroids
formed irregular, looser shapes (Figure 6B) that fragmented, particularly at higher drug
concentrations. The tumor morphology may depend on the competition between heterogeneous
cell proliferation caused by diffusion gradients of cell substrates, driving shape instability, and
stabilizing mechanical forces such as cell-cell and cell-matrix adhesion, as we quantitatively
showed in previous work [45]. Here, we observe a similar instability for drug-sensitive cells,
indicating that their adhesion is below the threshold necessary to maintain tumor shape stability.
This is confirmed through the computational model, where variation in the parameter values
for cell adhesion [22] reproduces these morphologies (Fig. 6C and D).
DISCUSSION
We have used an integrative methodology that tightly couples computational modeling with
biological data to quantify physiological resistance in 3-D tumor tissue in-vitro. The model
hypothesizes functional relationships that mechanistically link the growth and regression of a
tumor mass to the underlying phenotype. These relationships incorporate the complex interplay
between tumor growth, cell phenotype, and diffusion gradients, caused by heterogeneous
delivery of oxygen and cell substrates and removal of metabolites, and are thus able to calculate
tumor drug response as a predictable process dependent on biophysical laws. The results
underscore the importance of a quantitative approach in evaluating chemotherapeutic agents
that takes into consideration diffusion in addition to such chemo-protective effects as cell
phenotypic properties and cell-cell and cell-ECM (extracellular matrix) adhesion [49].
Although the model was calibrated with in-vitro data for breast cancer, this approach may also
apply to drug response of other tissues.
By modeling transport of drug and cell substrates, this work creates a quantitative tool that
could in the future be used to predict resistance in patients based on their tumor cell properties,
thus improving treatment outcome. In particular, the proposed functional relationships help to
quantify resistance in human breast cancer due to local cell substrate depletion. Our integrative
experimental/computational approach provides insight into the physical dynamics of solid
4256 nM exposure in-vitro for 96 hours yields an area-under-the-curve (AUC) of ~2.5×105 Molar.Hr, while for a typical patient Dox
plasma levels after a bolus injection represent an exposure of ~3×105 Molar.Hr.
5Median dose is the drug concentration required to achieve 50% cell kill.
Frieboes et al. Page 9
Cancer Res. Author manuscript; available in PMC 2010 May 15.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
tumors, and validates the hypothesized functional relationships as adequate to describe the
observed phenomena. Although more complex functional forms could conceivably also
provide a reasonable match between model and experimental results, quantifying the diffusion
effect with a minimal mathematical description is the preferred approach, unless it is known
to be wrong, as it facilitates insight into the system and provides for more economic simulations.
We explicitly chose not to use complex relationships that would contain more parameters than
the number of independent measurements available for calibration, thus resulting in an
underdetermined problem.
Cell packing density affects the magnitude of diffusion gradients and may pose a barrier to
effective drug penetration [27]. This density can vary between cell lines and tumors and reflect
variations in drug resistance (cf. Figure 6). How this relates to cell adhesion molecule (CAM)
expression (e.g., integrin, E-cadherin) is unclear. Mechanisms of cell packing may include
stronger cell adhesion forces, due to higher E-cadherin expression; this should also limit
proliferation, which does not seem to be the case here for the drug-resistant cells at lower drug
concentrations. Additionally, cellular stress affects the quantity and strength of CAMs. We
have previously investigated the effect of the tumor microenvironment on tissue morphology
[50,45], suggesting that marginally stable environmental conditions could directly affect
morphogenesis and present an additional challenge to therapy [48,45] in-vivo by increasing
tumor cell invasiveness and leading to complex infiltrative morphologies, depending on the
magnitude of cell adhesion forces that tend to maintain compact, non-invasive tumors [50].
Diffusion gradients combined with higher cell packing density augment drug resistance
synergistically, as observed in our experiments with the higher median dose for the more
compact, drug-resistant spheroids, yet it may be difficult to deterministically gauge their
combined impact [5]. Synergism may be due to increased drug binding to ECM in tumor areas
proximal to the drug source, while substrate and drug penetration to distal areas is significantly
reduced due to higher cell packing, thus exacerbating the resistance effect of the diffusion
barrier.
Note that the presumption that the necrotic region is not a risk factor for progression may not
be necessarily true because selection pressures for resistance and induction of mutation are
strongest in hypoxic, peri-necrotic areas. Thus, inducing a large necrotic fraction during
treatment may paradoxically further select for drug resistance.
The class of prediction modeling presented in this paper, based on an assimilation of complex
processes with a fully 3-dimensional physical environment, offers the capability of
complementing current pharmacokinetic measurements. A more expansive integration of
theoretical model parameters with biological data could help to move towards prediction of
treatment response based on in-vitro and in-vivo tumor information, and determine the
correspondence between in-vitro measurements and the in-vivo condition to refine and validate
the assumption that this approach can describe in-vivo tumors. Incorporation of patient-specific
tumor phenotypic and microenvironmental parameters into the model could enhance clinical
strategies and prognosis evaluation. We further envision that these methods will enhance
current pharmacokinetic models used in designing and interpreting Phase II clinical trials.
Supplementary Material
Refer to Web version on PubMed Central for supplementary material.
ACKNOWLEDGEMENTS
We acknowledge John Sinek and Steven Wise (Mathematics, UC-Irvine), Sandeep Sanga (Biomedical-Engineering,
UT-Austin), Paul Macklin (SHIS, UT-Houston), Ernest Han (Obstetrics/Gynecology, UC-Irvine), Hoa Nguyen
Frieboes et al. Page 10
Cancer Res. Author manuscript; available in PMC 2010 May 15.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
(Medicine, UC-Irvine) for helpful comments and discussions, and the reviewers for their valuable input. Grant support:
The Cullen Trust of Health Care, NSF-DMS 0818104, National Cancer Institute, Department of Defense.
REFERENCES
1. Trédan O, Galmarini CM, Patel K, Tannock IF. Drug resistance and the solid tumor microenvironment.
J Nat Cancer Inst 2007;99:1441–1454. [PubMed: 17895480]
2. Lankelma J, Dekker H, Luque RF, et al. Doxorubicin gradients in human breast cancer. Clin Cancer
Res 1999;5:1703–1707. [PubMed: 10430072]
3. Sinek J, Frieboes HB, Zheng X, Cristini V. Two-dimensional Chemotherapy Simulations Demonstrate
Fundamental Transport and Tumor Response Limitations Involving Nanoparticles. Biomed Microdev
2004;6:297–309.
4. Primeau AJ, Rendon A, Hedley D, Lilge L, Tannock IF. The distribution of the anticancer drug
doxorubicin in relation to blood vessels in solid tumors. Clin Cancer Res 2005;11:8782–8788.
[PubMed: 16361566]
5. Sinek JP, Sanga S, Zheng X, Frieboes HB, Ferrari M, Cristini V. Predicting drug pharmacokinetics
and effect in vascularized tumors using computer simulation. J Math Biol 2009;58:485–510. [PubMed:
18781304]
6. Greijer AE, de Jong MC, Scheffer GL, Shvarts A, van Diest PJ, van der Wall E. Hypoxia-induced
acidification causes mitoxantrone resistance not mediated by drug transporters in human breast cancer
cells. Cellular Oncology 2005;27:43–49. [PubMed: 15750206]
7. Spitz DR, Sim JE, Ridnour LA, Galoforo SS, Lee YJ. Glucose deprivation-induced oxidative stress in
human tumor cells. Annals NY Acad Sci 2000;899:349–362.
8. Lee YJ, Galoforo SS, Berns CM, et al. Glucose deprivation-induced cytotoxicity and alterations in
mitogen-activated protein kinase activation are mediated by oxidative stress in multidrug-resistant
human breast carcinoma cells. J Biol Chem 1998;273:5294–5299. [PubMed: 9478987]
9. Brown NS, Bicknell R. Hypoxia and oxidative stress in breast cancer. Oxidative stress: its effects on
the growth, metastatic potential and response to therapy of breast cancer. Breast Can Res 2001;3:323–
327.
10. Li J, Lee AS. Stress induction of GRP78/BiP and its role in cancer. Curr Mol Med 2006;6:45–54.
[PubMed: 16472112]
11. Tomida A, Tsuruo T. Drug resistance mediated by cellular stress response to the microenvironment
of solid tumors. Anti-Cancer Drug Design 1999;14:169–177. [PubMed: 10405643]
12. Yun J, Tomida A, Nagata K, Tsuruo T. Glucose-regulated stresses confer resistance to VP-16 in
human cancer cells through a decreased expression of DNA topoisomerase II. Oncol Res 1995;7:583–
590. [PubMed: 8704275]
13. Pusztai L, Hortobagyi GN. High-dose chemotherapy: how resistant is breast cancer? Drug Resist
Updat 1998;1:62–72. [PubMed: 17092798]
14. Sanga S, Sinek JP, Frieboes HB, Ferrari M, Fruehauf JP, Cristini V. Mathematical modeling of cancer
progression and response to chemotherapy. Expert Rev Anticancer Ther 2006;6:1361–1376.
[PubMed: 17069522]
15. Sanga S, Frieboes HB, Zheng X, Bearer E, Cristini V. Predictive oncology: a review of
multidisciplinary, multi-scale in-silico modeling linking phenotype, morphology and growth.
Neuroimage 2007;37:S120–S134. [PubMed: 17629503]
16. El-Kareh AW, Secomb TW. Two-mechanism peak concentration model for cellular
pharmacodynamics of Doxorubicin. Neoplasia 2005;7:705–713. [PubMed: 16026650]
17. Ward JP, King JR. Mathematical modeling of drug transport in tumour multicell spheroids and
monolayer cultures. Math Biosci 2003;181:177–207. [PubMed: 12445761]
18. Jackson TL. Intracellular accumulation and mechanism of action of doxorubicin in a spatiotemporal
tumor model. J Theor Biol 2003;220:201–213. [PubMed: 12468292]
19. Norris ES, King JR, Byrne HM. Modelling the response of spatially structured tumours to
chemotherapy: Drug kinetics. Math Comp Model 2006;43:820–837.
20. Byrne HM, Owen MR, Alarcón T, Murphy J, Maini PK. Modelling the response of vascular tumours
to chemotherapy: a multiscale approach. Math Models Meth Appl Sci 2005;16:1219–1241.
Frieboes et al. Page 11
Cancer Res. Author manuscript; available in PMC 2010 May 15.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
21. Panovska J, Byrne HM, Maini PK. A theoretical study of the response of vascular tumours to different
types of chemotherapy. Math Comp Model 2007;47:560–579.
22. Wise SM, Lowengrub JS, Frieboes HB, Cristini V. Nonlinear simulations of three-dimensional
multispecies tumor growth–I. Model and numerical method. J Theor Biol 2008;253:524–543.
[PubMed: 18485374]
23. Frieboes HB, Lowengrub JS, Wise S, et al. Computer simulation of glioma growth and morphology.
NeuroImage 2007;37:S59–S70. [PubMed: 17475515]
24. Zheng X, Wise S, Cristini V. Nonlinear simulation of tumor necrosis, neo-vascularization and tissue
invasion via an adaptive finite-element/level-set method. Bull Math Biol 2005;67:211–259.
[PubMed: 15710180]
25. Cristini V, Lowengrub J, Nie Q. Nonlinear simulation of tumor growth. J Math Biol 2003;46:191–
224. [PubMed: 12728333]
26. Sutherland RM. Cell and environment interactions in tumor microregions: the multicell spheroid
model. Science 1988;240:177–184. [PubMed: 2451290]
27. Grantab R, Sivananthan S, Tannock IF. The penetration of anticancer drugs through tumor tissue as
a function of cellular adhesion and packing density of tumor cells. Cancer Res 2006;66:1033–1039.
[PubMed: 16424039]
28. Fruehauf JP, Brem H, Brem S, et al. In vitro drug response and molecular markers associated with
drug resistance in malignant gliomas. Clin Cancer Res 2006;12:4523–4532. [PubMed: 16899598]
29. Fruehauf JP. In vitro assay-assisted treatment selection for women with breast or ovarian cancer.
Endocrine-Related Cancer 2002;9:171–182. [PubMed: 12237245]
30. Mehta RS, Bornstein R, Yu IR, et al. Breast cancer survival and in vitro tumor response in the Extreme
Drug Resistance Assay. Breast Cancer Res Treat 2001;66:225–237. [PubMed: 11510694]
31. Kern DH, Weisenthal LM. Highly specific prediction of antineoplastic drug resistance with an in
vitro assay using suprapharmacologic drug doses. J Nat Cancer Inst 1990;82:582–588. [PubMed:
2313735]
32. Roehm NW, Rodgers GH, Hatfield SM, Glasebrook AL. An improved colorimetric assay for cell
proliferation and viability utilizing the tetrazolium salt XTT. J Immunol Meth 1991;142:257–265.
33. Kim J, Kang K, Lowengrub J. Conservative multigrid methods for ternary Cahn-Hilliard systems.
Comm Math Sci 2004;2:53–77.
34. Jiang GS, Shu CW. Effcient implementation of weighted ENO schemes. J Comput Phys
1996;126:202–228.
35. Trottenberg, U.; Oosterlee, C.; Schüller, A. Multigrid. New York: Academic Press; 2001.
36. Isaacson, E.; Keller, H. Analysis of Numerical Methods. New York: Wiley; 1966.
37. Casciari JJ, Sotirchos SV, Sutherland RM. Glucose diffusivity in multicellular tumor spheroids.
Cancer Res 1988;48:3905–3909. [PubMed: 3383189]
38. Kallinowski F, Vaupel F, Runkel S, et al. Glucose uptake, lactate release, ketone body turnover,
metabolic micromilieu, and pH distributions in human breast cancer xenografts in nude rats. Cancer
Res 1988;48:7264–7272. [PubMed: 3191497]
39. Nugent LJ, Jain RK. Extravascular diffusion in normal and neoplastic tissues. Cancer Res
1984;44:238–244. [PubMed: 6197161]
40. Casciari JJ, Sotirchos SV, Sutherland RM. Variations in tumor cell growth rates and metabolism with
oxygen concentration, glucose concentration, and extracellular pH. J Cell Physiol 1992;151:386–
394. [PubMed: 1572910]
41. Teutsch HF, Goellner A, Mueller-Klieser W. Glucose levels and succinate and lactate dehydrogenase
activity in EMT6/Ro tumor spheroids. Eur J Cell Biol 1995;66:302–307. [PubMed: 7774614]
42. Acker H, Carlsson J, Mueller-Klieser W, Sutherland RW. Comparative pO2 measurements in cell
spheroids cultured with different techniques. Br J Cancer 1987;56:325–327. [PubMed: 3311111]
43. Durand RE. Slow penetration of anthracyclines into spheroids and tumors: a therapeutic advantage?
Cancer Chemother Pharmacol 1990;26:198–204. [PubMed: 2357767]
44. Kalra R, Jones A-M, Kirk J, Adams GE, Stratford IJ. The effect of hypoxia on acquired drug resistance
and response to epidermal growth factor in Chinese hamster lung fibroblasts and human breast-cancer
cells in vitro. Int. J Cancer 1993;54:650–655. [PubMed: 8514457]
Frieboes et al. Page 12
Cancer Res. Author manuscript; available in PMC 2010 May 15.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
45. Frieboes HB, Zheng X, Sun C-H, Tromberg B, Gatenby R, Cristini V. An integrated computational/
experimental model of tumor invasion. Cancer Res 2006;66:1567–1604.
46. Freyer JP, Sutherland RM. Regulation of growth saturation and development of necrosis in EMT6/
Ro multicellular spheroids by the glucose and oxygen supply. Cancer Res 1986;46:3504–3512.
[PubMed: 3708582]
47. Carlsson J. A proliferation gradient in three-dimensional colonies of cultured human glioma cells.
Int J Cancer 1977;20:129–136. [PubMed: 903181]
48. Gatenby RA, Smallbone K, Maini PK, et al. Cellular adaptations to hypoxia and acidosis during
somatic evolution of breast cancer. Br J Cancer 2007;97:646–653. [PubMed: 17687336]
49. Morin PJ. Drug resistance and the microenvironment: nature and nurture. Drug Resist Updates
2003;6:169–172.
50. Cristini V, Frieboes HB, Gatenby R, Caserta S, Ferrari M, Sinek J. Morphologic instability and cancer
invasion. Clin Cancer Res 2005;11:6772–6779. [PubMed: 16203763]
Frieboes et al. Page 13
Cancer Res. Author manuscript; available in PMC 2010 May 15.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Figure 1. Computational model parameters
Functional relationships linking tumor mass growth and regression to the underlying phenotype
are based on these parameters, with values derived from experimental observations (obtained
previously and in this study). (*) Value shown is an example, corresponding to drug-sensitive
MCF-7 WT cells at 256 nM Dox.
Frieboes et al. Page 14
Cancer Res. Author manuscript; available in PMC 2010 May 15.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Figure 2. Diffusion gradients are incorporated in the model functional relationships to simulate the
experimental conditions
Clockwise, top left: Immunohistochemical staining of tumor spheroids showing increasing
positivity for pimonidazole protein adducts (darker brown) in the direction of the center,
indicating a decrease in oxygen across the viable rim of ~ 100 µm (v.r.: viable rim; n.: necrotic
region); NHE-1 staining showing up-regulation of Na+/H+ transporters (darker brown) towards
the necrotic region in response to a decrease in pH; increasing positivity for GLUT-1 staining
(brown) in the peri-necrotic region demonstrating cellular adaptation to hypoglycaemic and
hypoxic stress due to decreasing concentration of glucose and oxygen across the spheroid
radius; increasing positivity of TUNEL staining (dark brown) towards the center indicating
correlation of apoptosis with hypoxia; Ki-67 nuclear staining (dark gray) showing cell
proliferation correlating with substrate availability. Bars, 50 µm.
Frieboes et al. Page 15
Cancer Res. Author manuscript; available in PMC 2010 May 15.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Figure 3. Computational model computes the gradients in 3-D tissue
Cross-section of a tumor spheroid in the computational model (A) (dotted line: tumor boundary;
dashed: perinecrotic area) showing calculated diffusion gradients (arrow) of cell substrate
concentration σ/σ (as fraction of external concentration in the medium) using diffusion
penetration length parameter L 100 µm. Necrosis becomes significant at substrate level σ/
σ < 0.5. (B) Doxorubicin concentration d/d in tumor tissue (normalized with concentration
in the medium) vs. distance from the tumor/medium interface, estimated for the model to be
consistent with data by [2,43], and others (Methods). By solving the unsteady Eq. (3), the drug
gradient across the viable region can be calculated at various times: bottom curve: 2 hours;
middle: 24 hours; topmost: steady state, showing ~ 2/3 drop from the tumor edge to the peri-
necrotic region.
Frieboes et al. Page 16
Cancer Res. Author manuscript; available in PMC 2010 May 15.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Figure 4. Cell proliferation and apoptosis
(A) Cell proliferation (ratio of proliferation-to-viability counts; n=3) in-vitro was lower by at
least 50% under conditions with diffusion gradients (spheroids) vs. monolayers for both cell
lines at drug concentrations d > 64 nM representing clinically relevant dosages. (B) Drug-
induced cell apoptosis death rate parameter (inverse time) vs. Doxorubicin
concentrations d calculated (Methods) from measured in-vitro monolayer viability counts
(n=3); glucose concentration σ = 2.0 g/L.
Frieboes et al. Page 17
Cancer Res. Author manuscript; available in PMC 2010 May 15.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Figure 5. Validation of the hypothesized functional relationships
This is done by the computational model quantifying the physiological resistance introduced
by the diffusion gradients. Cell viabilities as fraction of control N/NC vs. Doxorubicin
concentrations d; glucose concentration σ = 2.0 g/L; time T=96 hrs of drug exposure; n=3;
(A) MCF-7 WT drug-sensitive and (B) MCF-7 40F drug-resistant cells. White columns: 3-D
in-vitro tumor spheroids (dark gray: in-vitro monolayer data reported for comparison); light
gray: predictions of the computational model (error bars reflect variation in the apoptosis rate
parameter – see Methods).
Frieboes et al. Page 18
Cancer Res. Author manuscript; available in PMC 2010 May 15.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Figure 6. Correlation of tumor morphological stability with drug resistance
(A) Compact, nearly spherical morphology of a drug-resistant MCF-7 40F tumor spheroid in-
vitro is contrasted with the looser, irregular morphology of a drug-sensitive MCF-7 WT
spheroid (B). In the model, stable (C) and unstable (D) morphologies depend on variation in
the parameter values for cell adhesion forces [45]. Bar, 100 µm.
Frieboes et al. Page 19
Cancer Res. Author manuscript; available in PMC 2010 May 15.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
... 65 The remainder of this paper is structured as follows. We first elucidate our methods 66 by describing the underlying rules and parameters governing the cellular automaton as 67 well as the method of calculating oxygen transport and uptake. In the results section we 68 describe simulation results concerning healthy tissue growth in regularly arranged and 69 then heterogeneous vascular architectures. ...
... As each of these parameters has been estimated from different sources, we will fine 614 tune the basal oxygen consumption and physiologic vascular density for our specific 615 case using a well-studied tumour spheroid example. We utilize the observation that the 616 diffusion distance of oxygen to support cancer cells is approximately 10 cell 617 diameters [67] and the information from the literature concerning the ratio of cancer to 618 normal oxygen consumption. To estimate the baseline oxygen consumption rate then, 619 we begin with the value r c = 2.3 × 10 −16 mol cell −1 s −1 taken from an in vitro study of 620 tumour spheroid growth [51] and then perform a virtual tumour spheroid assay (Fig. 9) 621 to fine tune the value for our model system. ...
Preprint
Full-text available
Intratumoural heterogeneity is known to contribute to heterogeneity in therapeutic response. Variations in oxygen tension in particular have been correlated with changes in radiation response in vitro and at the clinical scale with overall survival. Heterogeneity at the microscopic scale in tumour blood vessel architecture has been described, and is one source of the underlying variations in oxygen tension. We endeavour to determine whether histologic scale measures of the erratic distribution of blood vessels within a tumour can be used to predict differing radiation response. Using a two-dimensional hybrid cellular automaton model of tumour growth, we evaluate the effect of vessel distribution on cell survival outcomes of simulated radiation therapy. Using the standard equations for the oxygen enhancement ratio for cell survival probability under differing oxygen tensions, we calculate average radiation effect in simulated, random, vessel organizations. We go on to quantify the vessel distribution heterogeneity and measure spatial organization using Ripley's L function, a measure designed to detect deviations from spatial homogeneity. We find that under differing regimes of vessel density the correlation coefficient between the measure of spatial organization and radiation effect changes sign. This provides not only a useful way to understand the differences seen in radiation effect for tissues based on vessel architecture, but also an alternate explanation for the vessel normalization hypothesis.
... Several cell population models exist in the literature, ranging from continuous to agent-based to hybrid models, and taking place at various scales (Schlüter et al. 2015;Fletcher et al. 2013;Szabó and Merks 2013;Deisboeck 2011;Cristini et al. 2003). Such models may reach a predictive power, where agreement/disagreement with biological data can advance our understanding of mechanistic relations within the biological systems (Jin 2016;Frieboes et al. 2009;Bearer 2009;Drasdo and Höhme 2005). Pertinent to the present work, previous research shows how analyzing the emergent morphology of cell population models can provide insight into the role of the model parameters (Giverso et al. 2016;Gerlee and Anderson 2007;Anderson et al. 2006), promoting future use of Bayesian methods. ...
Article
Full-text available
The landscape of computational modeling in cancer systems biology is diverse, offering a spectrum of models and frameworks, each with its own trade-offs and advantages. Ideally, models are meant to be useful in refining hypotheses, to sharpen experimental procedures and, in the longer run, even for applications in personalized medicine. One of the greatest challenges is to balance model realism and detail with experimental data to eventually produce useful data-driven models. We contribute to this quest by developing a transparent, highly parsimonious, first principle in silico model of a growing avascular tumor. We initially formulate the physiological considerations and the specific model within a stochastic cell-based framework. We next formulate a corresponding mean-field model using partial differential equations which is amenable to mathematical analysis. Despite a few notable differences between the two models, we are in this way able to successfully detail the impact of all parameters in the stability of the growth process and on the eventual tumor fate of the stochastic model. This facilitates the deduction of Bayesian priors for a given situation, but also provides important insights into the underlying mechanism of tumor growth and progression. Although the resulting model framework is relatively simple and transparent, it can still reproduce the full range of known emergent behavior. We identify a novel model instability arising from nutrient starvation and we also discuss additional insight concerning possible model additions and the effects of those. Thanks to the framework’s flexibility, such additions can be readily included whenever the relevant data become available.
... Mathematical modeling of breast cancer dynamics under treatment have gained interest for long time (Norton and Simon 1977;Enderling et al. 2006Enderling et al. , 2007Frieboes et al. 2009;Roe-Dale et al. 2011;Yankeelov et al. 2013;Lai et al. 2018;Jarrett et al. 2019;Lai et al. 2019Lai et al. , 2022. However, modeling of AI treatment in ER-positive breast cancer has received less attention so far. ...
Article
Full-text available
Estrogen receptor positive breast cancer is frequently treated with anti-hormonal treatment such as aromatase inhibitors (AI). Interestingly, a high body mass index has been shown to have a negative impact on AI efficacy, most likely due to disturbances in steroid metabolism and adipokine production. Here, we propose a mathematical model based on a system of ordinary differential equations to investigate the effect of high-fat diet on tumor growth. We inform the model with data from mouse experiments, where the animals are fed with high-fat or control (normal) diet. By incorporating AI treatment with drug resistance into the model and by solving optimal control problems we found differential responses for control and high-fat diet. To the best of our knowledge, this is the first attempt to model optimal anti-hormonal treatment for breast cancer in the presence of drug resistance. Our results underline the importance of considering high-fat diet and obesity as factors influencing clinical outcomes during anti-hormonal therapies in breast cancer patients.
... Mathematical modeling of breast cancer dynamics under treatment have gained interest for long time [20][21][22][23][24][25][26][27][28][29]. However, modeling of AI treatment in ERpositive breast cancer has received less attention so far. ...
Preprint
Full-text available
Estrogen receptor positive breast cancer is frequently treated with anti-hormonal treatment such as aromatase inhibitors (AI). Interestingly, a high body mass index has been shown to have a negative impact on AI efficacy, most likely due to disturbances in steroid metabolism and adipokine production. Here, we propose a mathematical model based on a system of ordinary differential equations to investigate the effect of high-fat diet on tumor growth. We inform the model with data from mouse experiments, where the animals are fed with high-fat or control (normal) diet. By incorporating AI treatment with drug resistance into the model and by solving optimal control problems we found differential responses for control and high-fat diet. To the best of our knowledge, this is the first attempt to model optimal anti-hormonal treatment for breast cancer in the presence of drug resistance. Our results underline the importance of considering high-fat diet and obesity as factors influencing clinical outcomes during anti-hormonal therapies in breast cancer patients.
... Friebiesm et al. [9] outline a PKPD model to quantify physiologic resistance of tumor tissue growing to drugs in a three-dimensional, in vitro case using a system of three second-order, non-linear PDEs that describe the spatio-temporal grows of the tumor and taking into consideration the interaction with the injected drug. The authors analyzed the in vitro case rather than the in vivo case in order to ignore the role of the metabolism in the dynamics, allowing the authors to calculate tumor drug response as a predictable process dependent on biophysical laws. ...
Preprint
Full-text available
Pharmaceutical nanoparticles (NPs) carrying molecular payloads are used for medical purposes such as diagnosis and medical treatment. They are designed to modify the pharmacokinetics-pharmacodynamics (PKPD) of their associated payloads, to obtain better clinical results. Currently, the research process of discovering the PKPD properties of new candidates for efficient clinical treatment is complicated and time-consuming. In silico experiments are known to be powerful tools for studying biological and clinical processes and therefore can significantly improve the process of developing new and optimizing current NPs-based drugs. However, the current PKPD models are limited by the number of parameters they can take into consideration and the ability to solve large-scale in vivo settings, thus providing relatively large errors in predicting treatment outcomes. In this study, we present a novel mathematical graph-based model for PKPD of NPs-based drugs. The proposed model is based on a population of NPs performing a directed walk on a graph describing the blood vessels and organs, taking into consideration the interactions between the NPs and their environment. In addition, we define a mechanism to perform different prediction queries on the proposed model to analyze two in vivo experiments with eight different NPs, done on mice, obtaining a fitting of 0.84 ± 0.01 and 0.66 ± 0.01 (mean ± standard deviation), respectively, comparing the in vivo values and the in silico results.
... In particular, breast tumors have been treated as porous media and fluid-like tissue. For example, Frieboes et al. modeled breast tumor tissues as porous media and implemented a model that incorporates the interplay between the local drug, oxygen, and nutrient concentrations [69]. On the other hand, Friedman and Hu argued that due to a high content of the extracellular fluid in mammary glands, breast tumors can be modeled by the Stokes equation [70], and their method has been successfully applied to many studies [71][72][73][74]. ...
Article
Full-text available
Theevolutionofbreasttumorsgreatlydependsontheinteractionnetworkamongdifferent cell types, including immune cells and cancer cells in the tumor. This study takes advantage of newly collected rich spatio-temporal mouse data to develop a data-driven mathematical model of breast tumors that considers cells’ location and key interactions in the tumor. The results show that cancer cells have a minor presence in the area with the most overall immune cells, and the number of activated immune cells in the tumor is depleted over time when there is no influx of immune cells. Interestingly, in the case of the influx of immune cells, the highest concentrations of both T cells and cancer cells are in the boundary of the tumor, as we use the Robin boundary condition to model the influx of immune cells. In other words, the influx of immune cells causes a dominant outward advection for cancer cells. We also investigate the effect of cells’ diffusion and immune cells’ influx rates in the dynamics of cells in the tumor micro-environment. Sensitivity analyses indicate that cancer cells and adipocytes’ diffusion rates are the most sensitive parameters, followed by influx and diffusion rates of cytotoxic T cells, implying that targeting them is a possible treatment strategy for breast cancer.
Chapter
In this chapter, a high level, namely tumour level, model and a corresponding simulation algorithm are introduced for the study of the effect of tumour-induced vessel displacement, on tumour progression rate prior to the onset of angiogenesis. The proposed model attributes its successful characteristics to a well known parallel bio-inspired computational tool, i.e. Cellular Automata (CAs) and solves diffusion differential equations to compute oxygen and glucose distribution into the tumour mass. CAs have been proven as efficient alternative to differential equations computational models that, despite their simplicity, exhibit complex dynamical behavior and can describe successfully the underlying phenomena for various physical, chemical and biological systems. More specifically, the studied phenomenon arrives from the displacement of the existing vessels, caused by mechanical forces owing to tumour growth, that heavily affect the aforementioned distribution of glucose and oxygen and, consequently and in turn, affect the tumour growth itself. The growth of a large number of tumours that initiated at various distances from a vessel, for the same number of time steps has been successfully modeled enabling us to further investigate and understand the underlying dynamics of early tumour growth. Simulation results showed that the properties of the physical diffusion processes in the case of moving vessel-tumour boundaries, affect directly tumour progression in the avascular progression phase. As such, the proposed model can be further utilized to explore various hypotheses of tumour growth relevant to drug delivery in chemotherapy, as well as to study access of growth factors and other plasma factors to the tumour.
Article
Immune cells in the tumor microenvironment (TME) are known to affect tumor growth, vascularization, and extracellular matrix (ECM) deposition. Marked interest in system-scale analysis of immune species interactions within the TME has encouraged progress in modeling tumor-immune interactions in silico. Due to the computational cost of simulating these intricate interactions, models have typically been constrained to representing a limited number of immune species. To expand the capability for system-scale analysis, this study develops a three-dimensional continuum mixture model of tumor-immune interactions to simulate multiple immune species in the TME. Building upon a recent distributed computing implementation that enables efficient solution of such mixture models, major immune species including monocytes, macrophages, natural killer cells, dendritic cells, neutrophils, myeloid-derived suppressor cells (MDSC), cytotoxic, helper, regulatory T-cells, and effector and regulatory B-cells and their interactions are represented in this novel implementation. Immune species extravasate from blood vasculature, undergo chemotaxis toward regions of high chemokine concentration, and influence the TME in proportion to locally defined levels of stimulation. The immune species contribute to the production of angiogenic and tumor growth factors, promotion of myofibroblast deposition of ECM, upregulation of angiogenesis, and elimination of living and dead tumor species. The results show that this modeling approach offers the capability for quantitative insight into the modulation of tumor growth by diverse immune-tumor interactions and immune-driven TME effects. In particular, MDSC-mediated effects on tumor-associated immune species' activation levels, average immune volume fraction, and influence on the TME are explored. Longer term, linking of the model parameters to particular patient tumor information could simulate cancer-specific immune responses and move toward a more comprehensive evaluation of immunotherapeutic strategies.
Article
Background: Breast cancer liver metastases (BCLM) are usually unresectable and difficult to treat with systemic chemotherapy. A major reason for chemotherapy failure is that BCLM are typically small, avascular nodules, with poor transport and fast washout of therapeutics from surrounding capillaries. We have previously shown that nanoalbumin-bound paclitaxel (nab-PTX) encapsulated in porous silicon multistage nanovectors (MSV) is preferentially taken up by tumor-associated macrophages (TAM) in the BCLM microenvironment. The TAM alter therapeutic transport characteristics and retain it in the tumor vicinity, increasing cytotoxicity. Computational modeling has shown that therapeutic regimens could be designed to eliminate single lesions. To evaluate clinically-relevant scenarios, this study develops a modeling framework to evaluate MSV-nab-PTX therapy targeting multiple BCLM. Methods: An experimental model of BCLM, splenic injection of breast cancer 4T1 cells was established in BALB/C mice. Livers were analyzed histologically to determine size and density of BCLM. The data were used to calibrate a 3D continuum mixture model solved via distributed computing to enable simulation of multiple BCLM. Overall tumor burden was analyzed as a function of metastases number and potential therapeutic regimens. Results: The computational model enables realistic 3D representation of metastatic tumor burden in the liver, with the capability to evaluate BCLM growth and therapy response for hundreds of lesions. With the given parameter set, the model projects that repeated MSV-nab-PTX treatment in intervals <7 days would control the tumor burden. Conclusion: Nanotherapy targeting TAM associated with BCLM may be evaluated and fine-tuned via 3D computational modeling that realistically simulates multiple metastases.
Article
Polymeric micelles are widely used as multifunctional drug carriers of poorly water‐soluble drugs, but the role of drug loading content is often overlooked. The purpose of this study was to investigate the cellular uptake and penetration of polymeric micelles with different drug loading contents and their effects on biological activities. In this study, poly(N‐(2‐hydroxypropyl) methacrylamide‐co‐methacrylic acid)‐block‐poly methyl methacrylate P(HPMA‐co‐MAA)‐b‐PMMA micelles were used as a nanocarrier for the encapsulation of the potent anticancer agent ellipticine (EPT). The micelles were loaded with various amounts of EPT and the physicochemical characteristics such as particle size, morphology and zeta potential of blank and EPT loaded nanoparticles were studied. Moreover, fluorescent lifetime studies confirmed that hydrophobic EPT is indeed in the PMMA micelle core. In vitro cytotoxicity tests using the glioma cell line U87MG revealed lower IC50 values when the cells were incubated with micelle with high drug loading content. The higher toxicity in micelles with higher drug loading content was associated with higher cellular uptake, which was monitored using laser scanning confocal microscopy and flow cytometry. Moreover, higher activity of micelles with higher drug loading was also observed in U87MG multicellular tumour spheroids although the difference was not significant. This article is protected by copyright. All rights reserved
Article
Full-text available
We previously observed that glucose deprivation induces cell death in multidrug-resistant human breast carcinoma cells (MCF-7/ADR). As a follow up we wished to test the hypothesis that metabolic oxidative stress was the causative process or at least the link between causative processes behind the cytotoxicity. In the studies described here, we demonstrate that mitogen-activated protein kinase (MAPK) was activated within 3 min of being in glucose-free medium and remained activated for 3 h. Glucose deprivation for 2–4 h also caused oxidative stress as evidenced by a 3-fold greater steady state concentration of oxidized glutathione and a 3-fold increase in pro-oxidant production. Glucose and glutamate treatment rapidly suppressed MAPK activation and rescued cells from cytotoxicity. Glutamate and the peroxide scavenger, pyruvate, rescued the cells from cell killing as well as suppressed pro-oxidant production. In addition the thiol antioxidant, N-acetyl-l-cysteine, rescued cells from glucose deprivation-induced cytotoxicity and suppressed MAPK activation. These results suggest that glucose deprivation-induced cytotoxicity and alterations in MAPK signal transduction are mediated by oxidative stress in MCF-7/ADR. These results also support the speculation that a common mechanism of glucose deprivation-induced cytotoxicity in mammalian cells may involve metabolic oxidative stress.
Article
Full-text available
We develop a conservative, second order accurate fully implicit discretization of ternary (three-phase) Cahn-Hilliard (CH) systems that has an associated discrete energy functional. This is an extension of our work for two-phase systems. We analyze and prove convergence of the scheme. To efficiently solve the discrete system at the implicit time-level, we use a nonlinear multigrid method. The resulting scheme is efficient, robust and there is at most a 1st order time step constraint for stability. We demonstrate convergence of our scheme numerically and we present several simulations of phase transitions in ternary systems.
Article
Full-text available
An existing multiscale model is extended to study the response of a vascularised tumour to treatment with chemotherapeutic drugs which target proliferating cells. The underlying hybrid cellular automaton model couples tissue-level processes (e.g. blood flow, vascular adaptation, oxygen and drug transport) with cellular and subcellular phenomena (e.g. competition for space, progress through the cell cycle, natural cell death and drug-induced cell kill and the expression of angiogenic factors). New simulations suggest that, in the absence of therapy, vascular adaptation induced by angiogenic factors can stimulate spatio-temporal oscillations in the tumour's composition. Numerical simulations are presented and show that, depending on the choice of model parameters, when a drug which kills proliferating cells is continuously infused through the vasculature, three cases may arise: the tumour is eliminated by the drug; the tumour continues to expand into the normal tissue; or, the tumour undergoes spatio-temporal oscillations, with regions of high vascular and tumour cell density alternating with regions of low vascular and tumour cell density. The implications of these results and possible directions for future research are also discussed.
Article
Empirical evidence and theoretical studies suggest that the phenotype, i.e., cellular- and molecular-scale dynamics, including proliferation rate and adhesiveness due to microenvironmental factors and gene expression that govern tumor growth and invasiveness, also determine gross tumor-scale morphology. It has been difficult to quantify the relative effect of these links on disease progression and prognosis using conventional clinical and experimental methods and observables. As a result, successful individualized treatment of highly malignant and invasive cancers, such as glioblastoma, via surgical resection and chemotherapy cannot be offered and outcomes are generally poor. What is needed is a deterministic, quantifiable method to enable understanding of the connections between phenotype and tumor morphology. Here, we critically assess advantages and disadvantages of recent computational modeling efforts (e.g., continuum, discrete, and cellular automata models) that have pursued this understanding. Based on this assessment, we review a multiscale, i.e., from the molecular to the gross tumor scale, mathematical and computational "first-principle" approach based on mass conservation and other physical laws, such as employed in reaction-diffusion systems. Model variables describe known characteristics of tumor behavior, and parameters and functional relationships across scales are informed from in vitro, in vivo and ex vivo biology. We review the feasibility of this methodology that, once coupled to tumor imaging and tumor biopsy or cell culture data, should enable prediction of tumor growth and therapy outcome through quantification of the relation between the underlying dynamics and morphological characteristics. In particular, morphologic stability analysis of this mathematical model reveals that tumor cell patterning at the tumor-host interface is regulated by cell proliferation, adhesion and other phenotypic characteristics: histopathology information of tumor boundary can be inputted to the mathematical model and used as a phenotype-diagnostic tool to predict collective and individual tumor cell invasion of surrounding tissue. This approach further provides a means to deterministically test effects of novel and hypothetical therapy strategies on tumor behavior.
Article
: Recently, glucose deprivation-induced oxidative stress has been shown to cause cytotoxicity, activation of signal transduction (i.e., ERK1, ERK2, JNK, and Lyn kinase), and increased expression of genes associated with malignancy (i.e., bFGF and c-Myc) in MCF-7/ADR human breast cancer cells. These results have led to the proposal that intracellular oxidation/reduction reactions involving hydroperoxides and thiols may provide a mechanistic link between metabolism, signal transduction, and gene expression in these human tumor cells. The current study shows that several other transformed human cell types appear to be more susceptible to glucose deprivation-induced cytotoxicity and oxidative stress than untransformed human cell types. In a matched pair of normal and SV40-transformed human fibroblasts the cytotoxic process is shown to be dependent upon ambient O2 concentration. A theoretical model to explain the results is presented and implications to unifying modern theories of cancer are discussed.
Article
Despite clear evidence that the effective penetration of the anthracycline antibiotics into experimental tumors or multicell spheroids is poor, these drugs exhibit clinical activity against a variety of solid tumors. In an attempt to understand this apparent contradiction, we used the Chinese hamster V79 spheroid system and flow cytometry techniques for intra-spheroid pharmacological studies of doxorubicin and daunomycin. Our results indicate that the slow delivery of the anthracyclines to the inner cells of spheroids is due to the rapid binding of the drug by cells in the outer layers. After exposure, the anthracyclines are retained much more effectively when cells remain in intact spheroids than when the spheroids have been dispersed, resulting in considerably more cytotoxicity in situ. This result indicates a need for considerable caution in attempting to predict the anti-tumor efficacy of drugs by using either conventional cell-culture systems, spheroids that have been disaggregated immediately post-exposure, or excision assays of tumors from experimental animals. Furthermore, our results suggest the need for a critical evaluation of the significance of the multidrug resistance (MDR) phenotype for cells surrounded by other drug-containing cells as opposed to single cells in drug-free culture medium.
Article
In this paper we consider the effects of a single anticancer agent on the growth of a solid tumour in the context of a simple mathematical model for the latter. The tumour is assumed to comprise a single cell population which reproduces and dies at a rate dependent on the local drug concentration. This causes cell movement and so establishes a velocity field within the tumour. We investigate the action of a single chemotherapeutic drug on the tumour and explore how different drug kinetics and treatment regimes may affect the final treatment outcome. A single infusion of drug is shown to be more effective than repeated short applications. We are able to construct asymptotic solutions to the model in the limit of a small drug degradation rate; these closely match solutions obtained numerically and provide additional insight into the behaviour of the tumour, in particular allowing the prediction of the strength of drug required to achieve tumour regression.
Article
In this paper we formulate and explore a mathematical model to study continuous infusion of a vascular tumour with isolated and combined blood-borne chemotherapies. The mathematical model comprises a system of nonlinear partial differential equations that describe the evolution of the healthy (host) cells, the tumour cells and the tumour vasculature, coupled with distribution of a generic angiogenic stimulant (TAF) and blood-borne oxygen. A novel aspect of our model is the presence of blood-borne chemotherapeutic drugs which target different aspects of tumour growth (cf. proliferating cells, the angiogenic stimulant or the tumour vasculature). We run exhaustive numerical simulations in order to compare vascular tumour growth before and following therapy. Our results suggest that continuous exposure to anti-proliferative drug will result in the vascular tumour being cleared, becoming growth-arrested or growing at a reduced rate, the outcome depending on the drug’s potency and its rate of uptake. When the angiogenic stimulant or the tumour vasculature are targeted by the therapy, tumour elimination can not occur: at best vascular growth is retarded and the tumour reverts to an avascular form. Application of a combined treatment that destroys the vasculature and the TAF, yields results that resemble those achieved following successful treatment with anti-TAF or anti-vascular therapy. In contrast, combining anti-proliferative therapy with anti-TAF or antivascular therapy can eliminate the vascular tumour. In conclusion, our results suggest that tumour growth and the time of tumour clearance are highly sensitive to the specific combinations of anti-proliferative, anti-TAF and anti-vascular drugs.