ArticlePDF Available

Investigation of solid tumor progression with account of proliferation/migration dichotomy via Darwinian mathematical model

Authors:

Abstract and Figures

A new continuous spatially-distributed model of solid tumor growth and progression is presented. The model explicitly accounts for mutations/epimutations of tumor cells which take place upon their division. The tumor grows in normal tissue and its progression is driven only by competition between populations of malignant cells for limited nutrient supply. Two reasons for the motion of tumor cells in space are taken into consideration, i.e., their intrinsic motility and convective fluxes, which arise due to proliferation of tumor cells. The model is applied to investigation of solid tumor progression under phenotypic alterations that inversely affect cell proliferation rate and cell motility by increasing the value of one of the parameters at the expense of another.It is demonstrated that the crucial feature that gives evolutionary advantage to a cell population is the speed of its intergrowth into surrounding normal tissue. Of note, increase in tumor intergrowth speed in not always associated with increase in motility of tumor cells. Depending on the parameters of functions, that describe phenotypic alterations, tumor cellular composition may evolve towards: (1) maximization of cell proliferation rate, (2) maximization of cell motility, (3) non-extremum values of cell proliferation rate and motility. Scenarios are found, where after initial tendency for maximization of cell proliferation rate, the direction of tumor progression sharply switches to maximization of cell motility, which is accompanied by decrease in total speed of tumor growth.
Content may be subject to copyright.
Journal of Mathematical Biology
https://doi.org/10.1007/s00285-019-01434-4
Mathematical Biology
Investigation of solid tumor progression with account
of proliferation/migration dichotomy via Darwinian
mathematical model
Maxim Kuznetsov1·Andrey Kolobov1,2
Received: 1 February 2019 / Revised: 21 June 2019
© Springer-Verlag GmbH Germany, part of Springer Nature 2019
Abstract
A new continuous spatially-distributed model of solid tumor growth and progression
is presented. The model explicitly accounts for mutations/epimutations of tumor cells
which take place upon their division. The tumor grows in normal tissue and its pro-
gression is driven only by competition between populations of malignant cells for
limited nutrient supply. Two reasons for the motion of tumor cells in space are taken
into consideration, i.e., their intrinsic motility and convective fluxes, which arise due
to proliferation of tumor cells. The model is applied to investigation of solid tumor
progression under phenotypic alterations that inversely affect cell proliferation rate
and cell motility by increasing the value of one of the parameters at the expense of
another.It is demonstrated that the crucial feature that gives evolutionary advantage
to a cell population is the speed of its intergrowth into surrounding normal tissue. Of
note, increase in tumor intergrowth speed in not always associated with increase in
motility of tumor cells. Depending on the parameters of functions, that describe phe-
notypic alterations, tumor cellular composition may evolve towards: (1) maximization
of cell proliferation rate, (2) maximization of cell motility, (3) non-extremum values
of cell proliferation rate and motility. Scenarios are found, where after initial tendency
for maximization of cell proliferation rate, the direction of tumor progression sharply
switches to maximization of cell motility, which is accompanied by decrease in total
speed of tumor growth.
Keywords Tumor progression ·Solid tumor growth ·Spatially distributed modeling ·
Intratumoral heterogeneity ·Tumor cell motility ·Mathematical oncology
Electronic supplementary material The online version of this article (https://doi.org/10.1007/s00285-
019-01434- 4) contains supplementary material, which is available to authorized users.
BMaxim Kuznetsov
kuznetsovmb@mail.ru
Extended author information available on the last page of the article
123
M. Kuznetsov, A. Kolobov
Mathematics Subject Classification 35Q92 ·92D15 ·92C17 ·35R09
1 Introduction
Our understanding of nature of cancer evolves with time, a lot of misleading theo-
ries having been produced throughout history (Sudhakar 2009). Although the correct
idea, that cancer arises due to genetic mutations, has been discussed already from the
beginning of the last century (Strong 1958), our knowledge about cancer is still very
far from being complete due to its incredible complexity. Nowadays, driven by the
widespread of oncological diseases, cancer research has become a very large inter-
disciplinary area. One of its relatively novel approaches is mathematical modeling of
tumor progression and tumor treatment, the popularity of which is now growing, not
least due to the rapid development of computer technology. Despite some skepticism,
expressed about mathematical modeling by oncologists, it has already proven to be
able to provide veridical insights into aspects of cancer growth (Boucher et al. 1990;
Gatenby et al. 2006) and to suggest effective optimizations of clinical protocols (Citron
et al. 2003; Chmielecki et al. 2011).
The modern concept of cancer was summarized at the end of the last century in
the famous article “The Hallmarks of Cancer” by Hanahan and Weinberg (2000). It
discusses six common capabilities, acquired by cancer cells. From the mathematical
point of view, four of the hallmarks can be combined into one concept, i.e., that cancer
cells can proliferate unlimitedly under sufficient level of nutrients (these hallmarks
are: limitless replicative potential, self-sufficiency in growth signals, insensitivity to
anti-growth signals and evading apoptosis). This was the crucial aspect already in the
first simplest models of tumor growth, which are governed by ordinary differential
equations and are still often used to fit experimental data (Skehan 1986;Hartetal.
1998). Account for the fifth hallmark—tissue invasion and metastasis—gave rise to
more complicated models, that also consider spatial distribution of cells. This is real-
ized via either continuous approach, i.e., using partial differential equations (Rockne
et al. 2009; Alfonso et al. 2016), or discrete approach, an example of which is cellu-
lar automaton (Jiao and Torquato 2011). The sixth hallmark, sustained angiogenesis,
i.e., formation of new blood vessels, was considered using different methods, includ-
ing complicated multiscale hybrid models (Macklin et al. 2009; Owen et al. 2009;
Kuznetsov et al. 2016).
The paradigm of cancer was expanded less than a decade ago by addition of four
hallmarks (Hanahan and Weinberg 2011), that further broadened the horizons for
mathematical modeling. New models of tumor growth can incorporate such features of
malignant tumors as deregulated metabolism (Kuznetsov and Kolobov 2018), evading
the immune system (Ledzewicz et al. 2012) and tumor-induced inflammation (Ander-
sen et al. 2017). However, probably the most fascinating hallmark for mathematical
modeling is the instability of cancer genome.
Nowadays it is known that each malignant tumor contains a large number of differ-
ent cell populations. This is partly due to very high rate of gene mutations in malignant
cells compared to that in normal ones (Bielas et al. 2006). Also, a comparable role in
intratumoral heterogeneity is played by epigenetic modifications, that result in herita-
123
Investigation of solid tumor progression with account...
ble alterations of cells phenotypes without changing their genotypes. The most studied
and probably the most important among such modifications are alterations in DNA
methylation and various modifications of histones (i.e., the proteins, that pack and
order DNA molecules) (Esteller 2008). Via these processes genes can be silenced or
activated. Frequent genetic and epigenetic alterations result in phenotypically hetero-
geneous and constantly changing cellular composition of a tumor. The progression
of disease is therefore determined by evolution of tumor cells under the influence of
selection, one of the driving factors of which is competition for nutritional resources
(Greaves and Maley 2012). The phenomenon of progression of cancer attracts wide
attention since it results in increase of its malignancy with time. Moreover, a great
challenge for cancer treatment is posed by the ability of cancer to acquire resistance to
drugs by propagation of this trait inside a tumor due to drug-induced selective pressure
(Gottesman 2002).
In mathematical modeling different approaches exist for consideration of intratu-
moral phenotypic heterogeneity and tumor progression. In continuous models, the
majority of which are not spatially distributed and thus are governed by ordinary dif-
ferential equations, the simplest approach is inclusion of the probability of transition
of a cancer cell upon its division into another predefined state, denoted by additional
variable, to which altered characteristics correspond. This approach is useful for con-
sideration of competition of small amount of phenotypically different populations,
the features of which differ drastically. The most popular feature, considered by this
method, is acquired drug resistance. Despite its simplicity, this approach allows to
address some of the major problems, frequently met in this kind of investigations,
such as estimating incidence of resistance to a drug under various parameters (Iwasa
et al. 2006) and proposing optimizations for scheduling of chemotherapy (Hadjian-
dreou and Mitsis 2014). In order to account for slow quantitative change of properties,
more complicated methods exist. One of them is to introduce a new variable for every
population which arises due to each mutation/epimutation, which happen upon division
of cells with some probability, and then to obtain new parameters for new variable via
some random alterations of the mother cell’s parameters (Stiehl et al. 2016). Another
method regards only one variable for description of tumor cells, but considers it as a
function of one or more specially introduced phenotypic parameters, which influence
characteristics of cells. Genome/epigenome instability can be expressed via diffu-
sion of cells in the space of phenotypic parameters, which implies an unphysiological
assumption that mutations can occur at any moment of cell cycle rather than only upon
DNA replication (Chisholm et al. 2015; Lorenzi et al. 2016). This can be overcome
by using integro-differential equation with integral kernel defining the probabilities of
various phenotypic shifts upon cell divisions (Lorz et al. 2013). Account for spatial
distribution significantly expands the spectrum of problems, that can be considered.
However, it is rarely met in continuous models, probably due to accompanying signif-
icant increase in computational cost. The existing works of this kind do not consider
mutations/epimutations of cells, multiple populations with various phenotypes being
already present from the beginning of simulations (Lorz et al. 2015; Lorenzi et al.
2018). However, spatial distribution is generally considered in the models, which sim-
ulate dynamics of tumor cells via discrete methods (Anderson et al. 2006; Gerlee and
Anderson 2008; Bouchnita et al. 2017). Such approach is very useful when dealing
123
M. Kuznetsov, A. Kolobov
with small amount of tumor cells, nevertheless, its common disadvantage is, as well,
great computational expense at consideration of later stages of tumor growth.
In this paper we present a new continuous spatially-distributed model of solid
tumor progression, comprised of a system of integro-partial differential equations,
with explicit consideration of mutations/epimutations of tumor cells. It is applied
to investigation of solid tumor progression with account of proliferation/migration
dichotomy, which represents the development of our previous work (Kolobov et al.
2000). Therein, in accordance with experimental studies by Theodorescu et al. (1991),
it has been demonstrated, that tumor can progress towards increase in cell motility
along with decrease in cell proliferation rate. This result is of merely qualitative nature,
and herein we investigate this phenomenon quantitatively. The prominent feature of
the presented model is the account for two reasons of cell motion, i.e., their intrinsic
motility and convective fluxes, which arise due to proliferation of tumor cells in dense
tissue.
2 Darwinian model of solid tumor progression
The model presented here consists of two separate blocks. The first block pro-
vides the mathematical description of changes in tumor cell composition due to
genetic/epigenetic mutations. It is based on the above-discussed use of integro-
differential equation accounting for phenotypic shifts upon divisions of cells, which
is further explained in detail. This block is integrated into the second one, which
represents a spatially-distributed model of growth of solid tumor in normal tissue.
2.1 Consideration of phenotypic instability
Let abe the general phenotypic parameter, which governs a set of characteristics
of malignant cells, that can be expressed numerically, including cell proliferation
rate B(a).Letn(a,t)denote the population density of tumor cells, so that a1
a0n(a,t)da
is the number of cells that possess values of the phenotypic parameter lying between a0
and a1at the moment of time t. We consider ato be scalar for simplicity, since
this is suitable for the purposes of this work. However, the following approach can
be straightforwardly expanded to the case of vector phenotypic parameter. Also, for
convenience, herein a∈[0,1].
We assume that, in a process of division of a cell with certain phenotypic parameter
value A, daughter cells can obtain altered values of it. The probability of obtaining
specific value of ais governed by a predefined function f(aA), which depends on the
difference between two parameter values. It is reasonable to assume, that this function
has a global maximum at aA=0, so that daughter cells most frequently receive the
values of the phenotypic parameter, close to A. The following function NewA(a,t)
describes the growth rate of tumor cells population density n(a,t)due only to the
emergence of offsprings of cells possessing phenotypic parameter value A,atthe
moment t:
123
Investigation of solid tumor progression with account...
NewA(a,t)=2B(A)n(A,t)·f(aA).
So far we assume, that cells of each population (i.e., which possess the same value
of phenotypic parameter) divide at constant rate, their proliferation being not limited
by nutrient depletion, overcrowding effect or any other reason. The factor 2 appears
in the formula due to the fact that upon a single cell division two new cells emerge,
phenotypes of which can be altered.
Figure 1illustrates the considered approach in a discrete manner, where Nis the
number of grid points in the permissible range of values for a. Thus, the following
description “converges” to themeaning of presented equations as N→∞. Solid line
depicts the graph NewA(a,t), which represents the distribution of new descendants
of population n(A,t), born over an infinitesimal period of time. The areas under this
curve, painted in different shades of gray, indicate the amounts of new cells of different
populations, that have appeared due to the proliferation of population n(A,t). Dashed
lines represent the distributions of daughter cells of other populations with a= A,
which may be transformed into one another, and into solid line, by shift along X-axis
and either stretch or compression along Y-axis, depending on the ratio of products of
cell proliferation rate and number of cells of two populations. In order to count, how
many new cells, which possess the value of a phenotypic trait A, emerge during this
time period, one needs to sum up the amounts of daughter cells of each population a,
that have undergone phenotypic shifts aA. They are depicted by overlapping red-
gray areas under the curves in the central grid region. In a strict mathematical form,
this means the following, where all functions are assumed to be integrable:
a(0,1)New(a,t)
1
0
NewA(a,t)dA=1
0
2B(A)n(A,t)·f(aA)dA.(1)
This concludes the essential part of the supposed approach. In this work, we use
normal distribution for the function, defining the probability of phenotypic shifts,
f(aA)=1
σ2πexp [aA]2
2σ2,(2)
and we apply the following boundary conditions, which mean that the cells, that are to
leave the permissible range of phenotypic parameter values, take its boundary values:
New(0,t)=
1
0
2B(a)n(a,t)·f(a)da
+
1
0
2B(a)n(a,t)·
a
−∞
f(z)dz
da,
123
M. Kuznetsov, A. Kolobov
2B(A)n(A,t)·f(1/N)
A-3/N A-2/N A-1/N A A+1/N A+2/N A+3/N
N
e
w
A
(a
,
t
)=
2
B
(
A
)
n
(A,
t
)·f
(
a
-
A
)
d
i
s
t
r
i
b
u
t
i
o
n
o
f
n
e
w
c
e
ll
s
a
m
o
u
n
t
o
f
n
e
w
c
e
l
l
s
~ B(a)n(a)
2B(A)n(A,t)·f(0)
values of phenotypic parameter a
Fig. 1 Illustration to the modeling approach for consideration of phenotypic instability of tumor cells. Here
a∈[0,1]is the general phenotypic parameter, its specific values are denoted below the x-axis, where Nis
the number of grid points; NewA(a,t)is the function of growth rate of tumor cells population density n(a,t)
due only to the emergence of new cells, mother cells of which possessed phenotypic parameter value A;
f(aA)is the function defining the probability of obtaining specific value of aby these new cells; B(a)is
cell proliferation rate. See the text for more explanations
New(1,t)=
1
0
2B(a)n(a,t)·f(1a)da
+
1
0
2B(a)n(a,t)·
1a
f(z)dz
da.(3)
If we had considered a non spatially-distributed case and had ignored cell death,
then the formulation of the problem would be completed by posing initial conditions,
defining the specific function for proliferation rates of different populations B(a)
and considering the following equation for overall change in tumor cells population
density, which accounts for emergence of new daughter cells and for “elimination” of
mother cells upon their division:
an(a,t)
t=New(a,t)B(a)n(a,t).
2.2 Tumor growth in tissue
Figure 2shows the block-scheme of interactions between variables in the model of
solid tumor growth in tissue. There are five variables in the model: density of tumor
123
Investigation of solid tumor progression with account...
n(a,r,t) – tumor cells
m(r,t) – necrosis
h(r,t) – normal cells
c(r,t) – capillaries
g(r,t) – glucose
m
cn h
g
Fig. 2 Block-scheme of the model of tumor growth in tissue. White arrows indicate transitions between
variables, green arrows denote stimulating effects that increase the value of a corresponding variable or the
intensity of a corresponding transition, red lines designate inhibitory interactions (color figure online)
cells n(a,r,t), density of normal cells h(r,t), fraction of necrosis m(r,t), concen-
tration of glucose g(r,t)and density of surface area of capillaries c(r,t). Space and
time coordinates rand tare further omitted for notational brevity.
The following set of equations governs the dynamics of tumor cells, normal cells
and necrosis:
Tumor cells: n(a)
t=
proliferation with redistribution

[New(a)B(a)n(a)]g
g+g
death

Mn(g)n(a)
migration

+D(an(a)
convection

−∇(In(a));
Normal cells: h
t=
death

Mh(g)h
convection

−∇(Ih);
Necrosis: m
t=
cell death

1
0
Mn(g)n(a)da+Mh(g)h
convection

−∇(Im);
where:
Mi(g)=1+tanh([gdg])
2Mmax
i,i=n,h;
1
0
n(a)da+m+h=1.
(4)
Herein two characteristics of cells are considered to vary between populations,
i.e., their proliferation rate B(a)and intrinsic motility D(a), the choice for which is
explained further in Sect. 3.1.
The term for proliferation of tumor cells takes into account the distribution of new-
born cells over phenotypes, controlled by Eqs. (1)–(3). Rate of tumor cells proliferation
falls under depletion of glucose. The form of the multiplier, which describes this effect,
implies, that proliferation rate of cells is proportional to their uptake of glucose, the
term for which will be discussed further. This assumption is due to the fact that glu-
cose is indispensable substrate for biosynthesis of several types of molecules, without
123
M. Kuznetsov, A. Kolobov
which cell division is impossible (Patra and Hay 2014). Since this multiplier depends
only on glucose level, it can be taken out of parentheses, leaving the function New(a)
unaltered.
Under a significant drop in glucose level, tumor cells, as well as normal ones, die
turning into necrosis. For description of cell death, a sigmoidal function Mi(g)is
used, which implies that cells begin to actively die under a critical level of glucose
in tissue. Tumor cells are known to obtain energy under lack of glucose via multiple
ways: catabolization of glutamine (Phan et al. 2014); utilization of lactate, generated
by other cells during aerobic glycolysis (Sonveaux et al. 2008); and many more,
including individual oxidative pathways for specific cell lines (Moreno-Sánchez et al.
2007). Therefore, the fact that in the model rate of cell death depends on the availability
of only one nutrient is an unavoidable simplification, justified by the fact that levels
of other nutrients would also drastically decrease along with that of glucose.
Tumor cells possess intrinsic motility and are able to migrate throughout the tissue,
which is described via diffusion-like term. We neglect the dependence of motility
of tumor cells on glucose uptake. Invasion of tumor cells into surrounding tissue is
one of the drivers of tumor growth which in existing mathematical models is often
considered as the only one (see for example Swanson et al. 2000; Alfonso et al.
2016). Such models are suitable for consideration of highly invasive tumors, like
glioblastoma. Herein we also consider another reason for tumor growth, i.e., pushing
surrounding tissues away by proliferating tumor cells. This type of growth is crucial to
low-grade compact tumors. During the tumor progression, its growth pattern generally
changes from compact to more invasive, therefore, the growth of real tumor often
represents a combination of these two types. Obviously, it is impossible to model the
growth of a non-invasive tumor via classical reaction–diffusion equations. Moreover,
neglect of convective transport will lead to underestimation of tumor growth speed.
Thus, simultaneous consideration of both types of solid tumor growth provides a
more physiologically correct modeling approach. This approach is realized herein by
inclusion of convective terms, which describe bulk motion of tissue elements.
The velocity field Iis determined by dynamics of tumor cells in tissue. To make its
derivation possible, it is necessary to introduce some additional constitutive assump-
tion. Herein it is the constancy of total density of cells and necrosis, which is normalized
to unity for convenience. This is expressed via the last equation in set of Eq. (4), where
for simplicity it is implied that the volumes of a normal cell and a tumor cell, as well
as of the amount of necrosis, formed in result of their death, are equal, while the volu-
metric fraction of capillary network in tissue in negligible. From the biophysical point
of view this assumption describes dense incompressible tissue, the former character-
istic meaning that total density of cells with necrosis cannot decrease, and the latter
that they cannot increase in any point of space in any moment of time. Equation for
velocity field Ican be obtained by summing up the equations for tumor cells, normal
cells and necrosis. During this procedure, left sides of equations produce the deriva-
tive of a constant value, i.e., zero; the transition terms of cell death are canceled out;
and the sum of convection terms provides the gradient of velocity field, which can be
expressed via remaining terms in the following way:
123
Investigation of solid tumor progression with account...
I=
cell proliferation

1
0
B(a)n(a)da·g
g+g
cell migration

+
1
0
D(an(a)da.
The final expression for Iwill have different forms for different space dimensions
and boundary conditions.
Of note, this approach for consideration of compact tumor growth represents a
special case of a more general method, utilized, e.g., by Byrne and Drasdo (2009).
That method converges to the one presented herein under the assumption that values
of solid pressure, arising due to cell dynamics, are infinitesimal, so they do not hinder
cell proliferation.
The equations for dynamics of glucose and capillaries are as follows:
Glucose: g
t=
inflow

Pc[gbl g]
consumption

−[Qn
1
0
n(a)da+Qhh]g
g+g
diffusion

+DgΔg.
Capillaries: c
t=
degradation

L[
1
0
n(a)da+m]c
convection

−∇Ic);
(5)
Dynamics of glucose is comprised from its inflow from capillaries into the tissue,
consumption by cells and diffusion throughout the tissue. The inflow of glucose is
governed by the process of passive diffusion through the walls of capillaries (Levick
2013). Therefore, the rate of glucose inflow is proportional to the density of capillaries’
surface area and to the difference in glucose concentrations in blood and in tissue.
Direct measurements demonstrate small arterial-venous blood glucose difference not
only in normal tissues, but even across tumors (see for example Eymontt et al. 1965),
so glucose concentration in blood is taken to be constant in the model.
Consumption of glucose by cells is described in traditional way, i.e., via equations
of Michaelis–Menten type. Tumor cells consume glucose much more actively than
normal ones, i.e., Qn>> Qh, since glucose is a crucial energy nutrient for tumor
cells and a major element for their proliferation (Vander Heiden et al. 2009). The rate
of glucose consumption does not vary for different populations of tumor cells, which
is discussed further in Sect. 3.1.
Capillary network degrades inside the tumor—it happens due to mechanical reasons
(Araujo and McElwain 2004), as well as due to various chemical factors (Holash
et al. 1999). Capillary degradation is not considered here in detail, and this process
is described by a rather phenomenological term. Like cells, capillaries move with the
convective fluxes, the speed of their motion being lower than that of cells due to their
connection with each other, i.e., γ<1. Other model parameters are provided and
justified in Sect. 3.2.
123
M. Kuznetsov, A. Kolobov
3 Investigation of tumor progression
3.1 Formulation of the task
The list of key properties that foster the prevalence of a cell population in the com-
position of tumor, obviously, includes its high proliferation rate—since it directly
governs the number of cells of this population—and its high intrinsic motility—since
this feature allows cells to leave nutrient-deficient regions inside main tumor mass and
promotes the spread of these cells into the surrounding tissues. As a rule, the degree
of tumor malignancy correlates with these indicators, since under the development of
a tumor, on a large time scale, more and more quickly dividing and rapidly migrating
cells are selected (see for example Louis et al. 2007).
However, for each individual tumor cell, its proliferation rate and motility are
inversely linked, i.e., the more a cell divides, the less it migrates, and vice versa
(Giese et al. 1996). This effect is, in particular, due to the fact that these processes
depend on two parallel metabolic pathways, and each of them relies on glucose: pen-
tose phosphate pathway generates necessary precursors for the synthesis of amino
acids, nucleotides and fatty acids; while glycolysis is the primary pathway for produc-
ing energy, utilized for motion of cancer cells (Shiraishi et al. 2015). It has been shown
by Kathagen-Buhmann et al. (2016) that inhibition of one of the enzymes, related to
either of these two processes, slows down the corresponding process and speeds up
another one (or in some cases at least does not affect the speed of another process
on a detectable level). This suggests that phenotypic alterations, which affect levels
of expression of corresponding enzymes, would lead to analogous effect. Thus, they
would have an ambiguous influence on the evolutionary adaptability of tumor cell
populations by enhancing one of the two crucial properties, that contribute to their
propagation in tumor composition, at the expense of another one.
In light of the above, herein we consider isolated phenotypic alterations of expres-
sion levels of virtual enzymes of either pentose phosphate pathway or glycolityc
pathway. A considered enzyme expression level in each case is denoted by pheno-
typic parameter aand is normalized in such a way, that a=0 refers to one of its
extremum values, corresponding to maximization of usage of glycolysis by a cell, and
thus to minimum cell proliferation rate and maximum motility that can be achieved
by variation of expression level of this enzyme. Vice versa, a=1 refers to maximum
proliferation rate and minimum motility of cells, while with the increase of aB(a)
monotonically increases and D(a)monotonically decreases. Of note, we assume, that
phenotypic alterations only redistribute the flow of glucose between the metabolic
pathways, not affecting the overall glucose consumption rate by cells. Alterations of
an enzyme expression are considered to be governed by normal distribution (see Eq. 2).
As it follows from the results of above mentioned experiments by Kathagen-Buhmann
et al. (2016), identical adjustments of expression levels of various enzymes lead to
quantitatively different changes in cell proliferation rate and motility, so we consider
various possible functions B(a), D(a)of the following general form:
123
Investigation of solid tumor progression with account...
B(a)=B0+[B1B0apB,
D(a)=D0+[D1D0apD,
where Bmin B0<B1Bmax,
Dmin D1<D0Dmax ,
0<pipmax,i=B,D,
(6)
where parameters B0,B1,pB,D0,D1,pDare varied from case to case.
The set of Eqs. (1)–(6) is the complete set of equations for the given study. Its main
question is what the general rules of tumor progression are under such formulation.
3.2 Parameters
The parameters of the model are assessed according to the data of various exper-
iments, if it is possible, and otherwise are estimated in order to reproduce known
characteristics of tumor growth. The basic set of parameters is given in Table 1, where
the following normalization parameters are used to obtain their dimensionless values:
tn=1 h for time, Ln=102cm for length, Sn=1 mg/ml for glucose concentration.
Normal capillary surface area density is equal to 100 cm2/cm3, which is its average
value for human muscle (Levick 2013). The value of glucose consumption rate for
human muscle (at rest) is also used in the model. Maximum density of tumor cells
is 3 ×108cells/ml, this value is taken from experimental work on growth of multicel-
lular tumor spheroids by Freyer and Sutherland (1985). Values for proliferation rate
of tumor cells and their glucose consumption rate are also estimated according to data
of this work. Coefficients of tumor cells motility correspond to non-invasive tumors,
being two to three orders of magnitude lower than the values for highly motile glioma
cells, estimated by Swanson et al. (2000). Maximum value of exponents for B(a)
and D(a)is adjusted in order to avoid sharp changes of the value of one parameter
under minor changes of the value of another one during phenotypic shift. Scale param-
eter for phenotypic instability is selected for the direction of tumor progression to be
apparent for almost all considered cases after already 2 to 300 days of tumor growth.
Maximum death rate of normal cells is much higher than that of tumor cells, since
the latter are known to be much less sensitive to metabolic stress (Izuishi et al. 2000).
Rather high value for capillaries degradation rate is dictated by results of the work on
high-resolution imaging of tumor microvasculature by Stamatelos et al. (2014), which
demonstrates that capillaries with adequate blood filling are already very scarce inside
the core of a tumor with the radius as small as 4 mm.
3.3 Numerical solving
For numerical solving the tumor was regarded as consisting of no more than
N+1=101 distinct cell populations n[A], where A=0,1/N,2/N,...,[N
1]/N,1, within each of which malignant cells are phenotypically identical. Before
the main computations a set of coefficients Fi,i=−N,N+1,...,1,0,1,...,
N1,Nwas introduced, which determine, what fraction of daughter cells of pop-
123
M. Kuznetsov, A. Kolobov
Table 1 Model parameters
Parameter Description Value Dimensionless value Estimations based on
Phenotypic instability of tumor cells
Bmin Minimum proliferation rate 0.01 h10.01 Freyer and Sutherland (1985)+seetext
Bmax Maximum proliferation rate 0.03 h10.03 Freyer and Sutherland (1985)+seetext
Dmin Minimum intrinsic motility 2.4×106cm2/day 0.001 Swanson et al. (2000)+seetext
Dmax Maximum intrinsic motility 2.4×105cm2/day 0.01 Swanson et al. (2000)+seetext
pmax Maximum exponent for B(a)and D(a)2.5 2.5 see text
σScale parameter for phenotypic instability 0.015 0.015 See text
Cell death
Mmax
nMaximum tumor cells death rate 0.002 h10.002 See text
Mmax
hMaximum normal cells death rate 0.02 h10.02 See text
Sensitivity to glucose level 5 5 See text
gdCritical level of glucose 0.56 mM 0.1 See text
Capillaries
LDegradation rate 4 ×1011 ml/cells s 0.012 Stamatelos et al. (2014)+seetext
γNetwork elasticity 0.5 0.5 See Sect. 2.2
123
Investigation of solid tumor progression with account...
Table 1 continued
Parameter Description Value Dimensionless value Estimations based on
Glucose
PPermeability of capillaries 1.1×105cm/s 4 Levick (2013)
gbl Blood level 5.56 mM 1 Association (2004)
QnTumor cells consumption rate 1.4×1016 mol/cell s 27 Freyer and Sutherland (1985)+seetext
QhNormal cells consumption rate 0.46 mg/min 100 ml 0.27 Baker and Mottram (1973)
gMichaelis constant for consumption rate 0.04 mM 0.007 Casciari et al. (1992)
DgDiffusion coefficient 2.6×106cm 2/s 94 Tuchin et al. (2001)
123
M. Kuznetsov, A. Kolobov
ulation A, that have emerged during a time step, belong to population A+i/N.
These coefficients were calculated as a set of definite integrals of function f(aA),
which governs the probabilities of phenotypic shifts, over the equal intervals of 1/N
(see Fig. 1), i.e.,:
F0=−er f 0.5
2σN,
FN=FN=1
21+er f 1
2σ,
i∈[1,N1]Fi=1
2er f i+0.5
22σNer f i0.5
2σN,
i∈[N+1,1]Fi=Fi.
The set of equations Eqs. (4)–(5), which governs the tumor growth, was solved in
one-dimensional region with a size of several centimeters. The exact size was selected
for each case in order to be sufficiently small to shorten computational time almost
as much as possible without imposing edge effects. Plane geometry was used for
computational simplicity, since it does not affect the results compared to spherically-
symmetrical case. Initial conditions corresponded to normal tissue with h(x,0)=1,
c(x,0)=1, m(x,0)=0, with a small, 0.1 mm in width, colony of tumor cells of
population with phenotypic parameter a=0.5 situated near the left boundary, where
thus n[0.5](x,0)=0.5, h(x,0)=0.5, c(x,0)=1. Other tumor cells were absent
in the beginning of a simulation: n[i](x,0)=0,i= 0.5. The initial distribution of
glucose g(x,0)was uniform and was calculated as its steady-state concentration in
normal tissue. For all variables zero-flux boundary conditions were set on the left
boundary; for glucose this condition was also set on the right boundary, while values
of other variables on the right boundary were constant, corresponding to normal tissue.
The convective flow speed was set to zero on the left boundary, and free boundary
condition was used for it on the right boundary. These conditions imply that the center
of the tumor is situated on the left boundary, and normal tissue, surrounding the
tumor, moves with constant speed outwards of the computational space through the
right boundary. Our model simulations may thus roughly represent, e.g., growth of
a cancer lump into a cavity, as it may happen in the case of leiomyoma. Boundary
conditions for convective flow speed result in the following equation for it:
I(x,t)=x
0
1
0
B(a)n(a,x,t)da·g(x,t)
g(x,t)+g
dr
+
1
0
D(an(a,x,t)da.(7)
Equation for glucose was considered in the quasi-stationary approximation due to
its fast dynamics and was solved using the tridiagonal matrix algorithm. For other
123
Investigation of solid tumor progression with account...
variables, the method of splitting into physical processes was used, i.e., kinetic equa-
tions, diffusion equations and convective equations were solved successively during
each time step. Kinetic equations were solved via second-order Runge–Kutta method,
Crank–Nicolson scheme was used for diffusion equations, and convective equations
were solved using the flux-corrected transport algorithm with explicit anti-diffusion
stage. The last method is introduced in the work by Boris and Book (1973), while
other classical methods are described in many books (see for example Press 2007).
The computational code was implemented in C++ and was run with spatial and tem-
poral discretisations sufficiently fine to not distort the solution.
3.4 Results
3.4.1 Different patterns of tumor progression
For the present study, a hundred sets of values of parameters B0,B1,D0,D1,pB,pD,
which define dependencies of cell proliferation rate and cell motility on phenotypic
parameter, B(a)and D(a), were generated in a random way, so that they satisfy
the conditions of Eq. (6). Then a hundred simulations of 500 days of tumor growth
were performed using these functions. For four of these sets, parametric plots, which
illustrate dependencies between Band D,areshowninFig.3a. For these cases,
Fig. 3b demonstrates the change in time of the average value of phenotypic parameter
throughout all the tumor cells, a, which is calculated as follows:
a(t)=
L
0
1
0
a·n(a,x,t)dadx/
L
0
1
0
n(a,x,t)dadx
As Fig. 3b shows, there is no single direction of tumor progression in terms of
maximizing either proliferation rate or motility of cells. In case 1, from the beginning
of tumor growth, more and more rapidly proliferating populations displace the others
until aapproaches the value, close to unity, overcoming of which is restrained due to
boundary conditions (see Eq. (3)). In case 2, the composition of tumor changes so that
more and more motile populations succeed. Case 3 demonstrates slow change of a,
which resides near its initial value. In this case, distribution of number of tumor cells
over phenotypic parameter n(a,t)=L
0n(a,x,t)dxremainstobesinglemode(see
inlet). Thus, the cells, which possess maximum of either proliferation rate or motility,
turn to be less fit in this case, than the cells with phenotypic value, close to 0.5. In
case 4, the direction of tumor progression, which initially inclines towards increase
in cell proliferation rate, changes after 187 days, eventually leading to domination of
more motile cells. Unlike the previous case, here during the switch in direction of
progression, the distribution n(a)is bimodal. It has maxima at two boundary values
of a, so that the value of aduring this switch is determined by the ratio of numbers of
cells, which possess values of phenotypic parameter, close to the boundary ones.
123
M. Kuznetsov, A. Kolobov
0
0.25
0.5
0.75
1
B,x10
-2
D,x10-3
0 100 200 300 400 500246810
1
1.5
2
2.5
3
(a) (b)
1
2
4
3
days
a
1
2
4
3
a=1
a=0.5
a=0
х
х
х
proliferation rate
increases
motility
increases
a
relative number
of cells
01
a
relative number
of cells
01
Fig. 3 aDependencies between cell proliferation rate Band intrinsic cell motility Dduring phenotypic
shift in four considered cases, where the following functions are used: (1) B(a)=0.014 +0.016 ·a2,
D(a)=0.006 0.002 ·a1.3;(2)B(a)=0.015 +0.003 ·a2.1,D(a)=0.008 0.005 ·a0.6;
(3) B(a)=0.015 +0.013 ·a1.5,D(a)=0.009 0.007 ·a1.8;(4)B(a)=0.012 +0.013 ·a2.3,
D(a)=0.009 0.006 ·a0.4.bChange of average phenotypic parameter over time during simula-
tion of tumor growth for four cases, shown in (a). Crosses denote the moments, for which distribution of
model variables for case 4 are shown in Fig. 4
3.4.2 Visualization of tumor growth
Figure 4provides snapshots of spatial distributions of model variables for case 4.
In the beginning of the simulation small amount of tumor cells actively proliferate
under abundance of glucose near the left boundary, which represents the center of a
tumor. Since during this period spatial effects play insignificant role, tumor cells, that
proliferate faster, always hold evolutionary advantage in the very beginning of tumor
growth. Increase in number of tumor cells results in significant local degradation of
capillaries, followed by decrease in glucose inflow. Deficiency of glucose leads to
formation of necrosis in tumor center, while proliferating cells concentrate on the
outer region, which possesses sufficient amount of glucose for cell division. Prolif-
erating cells push away the cells, that are located further from the tumor center, that
drives tumor front propagation. Moreover, tumor cells, that migrate towards region
with normal microvasculature density, obtain more glucose, which stimulates their
proliferation and further migration away from tumor center. That represents another
reason for propagation of tumor front. We will refer to the first process as “convective
growth” and to the second process as “intergrowth”.
After the initial period of tumor growth, spatial effects become determinative for
tumor progression. Limited nutrient supply poses selective pressure on tumor cells and
phenotype selection is driven by competition for this supply. Cells that are able to reach
glucose-abundant region, proliferate, while cells in the tumor core die out. As it was
noticed above, this competition may result in different patterns of tumor progression,
and its direction may even change during tumor growth, like in case 4. Figure 4b refers
123
Investigation of solid tumor progression with account...
26 31
30292827
1.0
0.2
0.0
0.4
0.6
0.8
distance to tumor center, mm
day 380
capillaries, c
glucose, g
tumor cells
with necrosis, n+m
1.0
0.2
0.0
0.4
0.6
0.8
3021
distance to tumor center, mm
day 6
1.0
0.2
0.0
0.4
0.6
0.8
10 15
14131211
tumor cells, n
distance to tumor center, mm
day 180
(a) (b)
(c)
Fig. 4 Profiles of tumor cells density n, necrosis fraction m, microcirculatory network surface area density c
and glucose concentration gunder B(a)=0.012 +0.013 ·a2.3,D(a)=0.009 0.006 ·a0.4(case 4 on
Fig. 3)ona6-th day of tumor growth, b180-th day of tumor growth, c380-th day of tumor growth
to the 180-th day of tumor growth for this case, which is at this moment accompanied
by dominance of more rapidly proliferating tumor cells. Figure 4c demonstrates the
380-th day of simulation, until which more motile cells have already displaced rapidly
proliferating ones in tumor composition, which resulted in reduction in maximum
density of tumor cells, more gentle slope of tumor front and slightly deeper penetration
of tumor cells in adjacent normal tissue.
3.4.3 Reasons for evolutionary advantage of tumor cells
The provided description of tumor growth dynamics allows to suggest that cells, which
are able to migrate away from the main tumor mass and reach nutrient-abundant region,
should be evolutionary advantageous. Therefore, it is the speed of their intergrowth
into normal tissue, that should play crucial role for the evolutionary fitness of a cell
population—since in every point in space convective fluxes shift all the cells with the
same speed, while the speed of intergrowth may vary for different populations. The
intergrowth speed should be an increasing function of both cell proliferation rate Band
motility D. This idea is favored by the fact, that for the well-known Fisher’s equation,
n
t=Bn(1n)+Dn
xx, which can be regarded as a simple model of tumor growth via
123
M. Kuznetsov, A. Kolobov
only the process of intergrowth, the formula for the wave speed is V2BD (Fisher
1937). Derivation of a similar analytical formula for the set of equations, considered
herein, is a hardly feasible task due to its complexity, i.e., inclusion of convective
terms, interplay between tumor cell populations and consideration of another variables.
However, intergrowth speeds of cell populations can be estimated numerically via the
following approach.
Figure 5a demonstrates the map of tumor front propagation speeds for monoclonal
tumors, the cells of which possess values of Band Din the ranges Bmi n BBmax,
0DDmax. The total speed of tumor front can be interpreted as the sum of two
speeds, related separately to convective growth and intergrowth. This is an approxi-
mation, valid only for small enough values of cell motility. Generally, these processes
are not independent, due to the fact that the process of intergrowth redistributes pro-
liferating cells and affects convective fluxes. We define convective speed for a single
cell population as the speed of propagation of a monoclonal tumor with the same
cell proliferation rate and zero cell motility. Further, we estimate intergrowth speed
of a cell population as the total speed of such monoclonal tumor front propagation
minus its convective speed. The map of intergrowth speeds, obtained via this method
for monoclonal tumors is provided in Fig. 5b, where the plots B(D)are repeated for
four considered cases. Note that intergrowth speed on average makes up almost 50%
of total speed of tumor growth within considered range of parameters. Obviously,
in a polyclonal tumor intergrowth speed of a single population may be affected by
simultaneous dynamics of other populations (this effect will be demonstrated further
in Sect. 3.4.4). Nevertheless, intergrowth speed, assessed via the described approach,
serves as a good indicator for direction of tumor progression.
Figure 5c provides the duplication of graphs of average phenotypic parameter by
time for considered cases of tumor progression, along with the heat maps of dependen-
cies of intergrowth speeds of corresponding populations on the value of phenotypic
parameter a. This figure demonstrates, that for cases 1, 2 and 4, in which intergrowth
speed reaches its maximum at one of the boundary values of a, tumor indeed progresses
towards maximization of its overall intergrowth rate, reaching the near-extreme values
of a, overcoming of which is restricted due to boundary conditions. However, in case 3
the value of anoticeably exceeds the value of a, which corresponds to the population,
possessing maximum intergrowth speed.
The qualitative behaviour of aover time for all other simulations in each case
corresponds to one of the four presented cases. The figures, analogous to Fig. 5b, c
for all the cases are provided as Supplementary files. Among 100 cases, in 17 of them
tumor progresses towards maximization of cell proliferation rate, like in case 1; in 51
cases tumor progresses towards maximization of cell motility, among which in 45
cases this direction of evolution is chosen from the early stage of tumor growth, like
in case 2, and in 6 cases the direction of evolution undergoes a distinguishable switch,
like in case 4; and finally, in 32 cases tumor progresses towards non-extremum values
of cell proliferation rate and motility, like in case 3. In all cases of third type the value
of average phenotypic parameter also exceeds the value, which corresponds to the
population with the maximum intergrowth speed.
123
Investigation of solid tumor progression with account...
B,x10-2
D,x10-3
1
1.5
2
2.5
3
B,x10-2
D,x10-3
1
1.5
2
2.5
3
(b)
mm/week
0
1
2
4
3
a=1
a=0.5
a=0
0
06810024246810
100 200 300 400 500
0
0.25
0.5
0.75
1
(c) days
a
1
2
4
3
0.321
0.323
0.346
0.380
0.407
Total speed Intergrowth speed
0.404
0.332
0.293
0.258
0.216
0.435
0.450
0.452
0.400
0.181
0.373
0.283
0.269
0.268
0.263
0.455
0
0.25
0.5
0.75
1
a
134
2
Intergrowth speed
proliferation rate
increases
motility
increases
0.2
0.4
0.6
0.8
1.0
1.2
1.4
(a)
Fig. 5 aSpeed of monoclonal tumor growth in dependence of values of its cells’ proliferation rate B
and intrinsic motility D.bEstimated speed of intergrowth of tumor population in surrounding tissue in
dependence of values of Band D. Lines 1–4 designate the considered cases of behaviour of functions B(a)
and D(a)during phenotypic shift, and match the lines, presented in Fig. 3a. cGraph of average phenotypic
parameter over time for four considered cases, which matches the graph presented in Fig. 3b, along with
the heat maps of intergrowth speed of populations for these cases. Bold labels denote maximum values of
intergrowth speed for each case
3.4.4 Under close intergrowth speeds, populations with higher proliferation rate
prevail in tumor composition
The fact that cell populations with higher proliferation rate may prevail in tumor
composition despite lower intergrowth speed, can be explained via consideration of
tumor cell dynamics in case 3. After initial period of tumor growth, a whole spectrum
of cell populations with noticeably different proliferation rates and motilities form a
complex phenotypic structure, illustrated in Fig. 6, where local average phenotypic
value, ˆa, is calculated in the following way:
123
M. Kuznetsov, A. Kolobov
capillaries, c
glucose, g
tumor cells
with necrosis, n
+m
tumor cells, n
0.4
0.3
0.2
0.1
53.0 53.4 53.8 54.2
distance to tumor center, mm
0.418
0.437
0.456
0.475
0.494
0.380
0.399
a
Fig. 6 Distribution of populations along with profiles of model variables at the tumor rim on the 500-th day
of tumor growth under B(a)=0.015 +0.013 ·a1.5,D(a)=0.009 0.007 ·a1.8(case 3 in Figs. 3,5).
Variation of average phenotypic parameter in space is denoted with different shades of blue (color figure
online)
ˆa(x,t)=
1
0
a·n(a,x,t)da/
1
0
n(a,x,t)da.
The qualitative appearance of this structure persists during tumor growth, accom-
panied by only a very slight increase in aat least until the end of simulation. Such
kind of structure is maintained due to complex interplay between populations of tumor
cells. Distribution of cells with a=0.455, which possess higher nominal intergrowth
speed, has its maximum at the distance of 54 mm from the tumor center. There
they actively divide, producing cells of populations with both equal, lower and higher
values of a. Cells with its lower values have higher motility and lower proliferation
rate. However, some of these cells, which due to the process of random migration
have moved several millimeters forward, obtain more glucose and therefore are able
to proliferate faster, that effectively increases their intergrowth speed and stabilizes
their position in front of cells with a=0.455. Nevertheless, the majority of emerging
new cells with a= 0.455, with lower nominal intergrowth speeds, move away from
the outer part of tumor. Since the cells, that possess higher value of a, proliferate faster
under the same levels of glucose, the value of ˆaincreases towards the tumor center.
Thus, the majority of tumor cells correspond to values of a>0.455.
3.4.5 Speed of tumor growth may fall in result of competition between cells
Cases 3 and 4 demonstrate, that tumor progression does not always result in max-
imization of total tumor growth speed. In case 3 cell population with a0.7 has
maximum total speed, while tumor progresses towards lower average phenotypic
parameter a0.5. In case 4 cell population with a=1 has maximum total speed,
and though in the beginning of tumor progression ainclines towards it, eventually it
switches towards the lowest values of a.
The dynamics of tumor composition for case 4 in shown in Fig. 7, where the average
value of phenotypic parameter ˆais marked for the points, where tumor cell density
123
Investigation of solid tumor progression with account...
g=g*=0.007
(proliferation rate falls two times)
400
0
100
200
300
025
20
0151
530
days
distance to tumor center, mm
0.0
0.2
0.4
0.6
0.8
a
n=0.025
n=0.25
for glucose:
Isolines
for tumor cells:
posion that tumor front would have
without fall in speed of its propagaon
Fig. 7 Distribution of populations during first 400 days of tumor growth under B(a)=0.012 +0.013 ·a2.3,
D(a)=0.009 0.006 ·a0.4(case 4 on Figs. 3,5). Variation of average phenotypic parameter in space and
time is denoted with different shades of blue. Isolines for indicated values of tumor cell density and glucose
concentration are shown (color figure online)
exceeds 106. The graph is arranged so that horizontal slices provide a distribution
of average local phenotypes within the tumor at certain moments of time. As Fig. 5c
indicates, for this case the variation of intergrowth speed of cell populations is rather
moderate in the range from a0.4toa=1, which initially drives progression of
tumor towards increase in cell proliferation rate. However, as the average phenotypic
parameter within the main tumor mass increases, a small region at the tumor rim
becomes occupied by cells, which have low values of a. Such cells emerge rarely at
the beginning of tumor growth, however, since their populations have significantly
higher intergrowth speed, they occupy more and more space in glucose-rich tumor
rim, and eventually cut off the nutrient supply for the cells, which have high values
of a, further replacing them within the main tumor mass. Of major interest is the fact
that the speed of tumor front propagation decreases during this process by 15%, i.e.,
from 0.62 mm/week at the 180-th day to 0.53 mm/week at the 380-th day. It happens
due to the fact, that remaining tumor cell populations possess lower proliferation rate
(which in this case differs about twice-fold under a=0 and a=1), which results
is significantly lower convective flow speed. Such decline in tumor front propaga-
tion is manifested in 5 out of 6 performed simulations where the direction of tumor
progression undergoes analogous distinguishable switch.
4 Discussion
The results of investigation of solid tumor progression under isolated phenotypic alter-
ations, that inversely affect proliferation rate and motility of cells, were presented.
123
M. Kuznetsov, A. Kolobov
Tumor progression was considered to be driven only by competition between popula-
tions of malignant cells for limited nutrient supply. Two types of tumor growth were
considered simultaneously, i.e., convective growth and intergrowth into surrounding
normal tissue. It was shown, that the crucial feature that renders a cell population evo-
lutionary advantageous is its intergrowth speed. Four qualitatively different patterns
of tumor progression were found, which show that the increase in intergrowth speed
of a tumor, which always accompanies its progression, is not always associated with
increase in motility of its cells. Scenarios were found, in which tumor progression is
accompanied by decrease in total speed of tumor growth.
As we have stated in Sect. 3.4.3, two considered types of growth are not indepen-
dent, since under high motility of cells the process of their intergrowth into surrounding
tissue redistributes the profile of proliferating cells and affects convective fluxes. An
investigation of the interaction of two types of tumor growth is a fascinating task.
However, consideration of growth of highly diffusive tumors using the presented sub-
model of tumor growth under the same basic values of parameters would lead to
unphysiologically explosive tumor growth. Such limitation is associated with a more
general problem of modeling microvasculature network via spatially-distributed vari-
ables, that is unavoidably phenomenological. In reality a small but sufficient local
aggregation of tumor cells will already lead to their competition for nutrients, since
the deeper situated cells will lack them. On the contrary, in a continuous model the
cells are free to grow until rather high numbers of them, because all the variables in
every computational grid are effectively mixed. Moreover, in reality degradation of a
capillary leads to the cessation of downstream blood flow and thus effective shutdown
of the whole branch of capillaries. In case of highly invasive tumors this effect would
be of great importance, however, it is impossible to straightforwardly incorporate it
into the utilized continuous approach.
Explicit consideration of mutations/epimutations of tumor cells upon their divi-
sion results in a system of integro-partial differential equations, numerical solution of
which is rather computationally expensive. In the area of modeling of tumor progres-
sion, there exist at least two more simple approaches, already mentioned in Sect. 1,
which might be used for the considered task, reducing its numerical cost. One of them
is consideration of heterogeneous tumor composition at initial moment and neglect
of mutations/epimutations of cells. However, preliminary simulations utilizing this
approach showed that it is not suitable for the considered task, since it produced qual-
itatively different results depending on the initial phenotypic composition of tumor.
Another method is consideration of diffusion of cells in the space of phenotypic param-
eters, which accounts for cell cycle-independent phenotypic alterations. Preliminary
simulations with this method showed good qualitative correlation with the results
presented herein. Nevertheless, such correlation may fail in cases when significant
alterations in cell proliferation rate accompany tumor growth. An example of such
situation is consideration of antiangiogenic therapy, which reduces the inflow of nutri-
ents to the tumor and thus affects proliferation rate and influences the frequency of
phenotypic shifts. This led to the choice of the utilized approach for the given study,
which may serve as an approbation for its further usage.
It should be noted, that it is an open question, whether some heritable epigenetic
modifications can happen independently of cell division, certain data suggesting this
123
Investigation of solid tumor progression with account...
possibility. For example, methylation of DNA is known to be controlled by a family
of enzymes, a part of which is aimed to maintain DNA methylation pattern upon its
replication—thus, epimutations, produced by its errors, certainly happen upon cell
division; however, a part of enzymes perform de novo methylation, and the details
of this process are still not elucidated completely (Robertson 2001). Nevertheless,
certain experimental findings by Velicescu et al. (2002) suggest that cell division is
nevertheless required for de novo methylation, which also speaks in favor of the chosen
approach.
The presented mathematical model is, to the best of our knowledge, the first continu-
ous spatially-distributed model of solid tumor progression, which explicitly considers
mutations/epimutations of tumor cells. It has been used herein for solution of only
one particular problem, and it can serve as a framework for the following studies
focusing on different phenomena of cancer progression. There are a lot of ways to
upgrade the model for such tasks, moreover, the choice of a character of phenotypic
changes, governed by alternating phenotypic parameter, is limited only by the imag-
ination of the researcher. One valuable option is introducing the dependency of the
scale parameter for phenotypic instability σon the phenotypic parameter a, which
accounts for the fact that probability of a mutation upon a single replication of a cell
increases with increase in its malignancy. The submodel of tumor growth in tissue
can be also upgraded, and more complicated versions of it have already been used
in our previous papers (Kolobov and Kuznetsov 2015; Kuznetsov and Kolobov 2017,
2018; Kuznetsov et al. 2018). One of the intriguing options, that we are planning to
implement, is modeling of chemotherapy, that will pose additional selective pressure
on tumor cells.
Acknowledgements The reported study was funded by RFBR according to the research Projects Nos.
16-01-00709, 17-01-00070 and 19-01-00768. Numerical simulations have been prepared with the support
of the “RUDN University Program 5-100”.
References
Alfonso J, Köhn-Luque A, Stylianopoulos T, Feuerhake F, Deutsch A, Hatzikirou H (2016) Why one-size-
fits-all vaso-modulatory interventions fail to control glioma invasion: in silico insights. Sci Rep UK
6:37283
Andersen M, Sajid Z, Pedersen R, Gudmand-Hoeyer J, Ellervik C, Skov V, Kjær L, Pallisgaard N, Kruse
T, Thomassen M et al (2017) Mathematical modelling as a proof of concept for MPNS as a human
inflammation model for cancer development. PLoS ONE 12(8):e0183620
Anderson A, Weaver A, Cummings P, Quaranta V (2006) Tumor morphology and phenotypic evolution
driven by selective pressure from the microenvironment. Cell 127(5):905–915
Araujo R, McElwain D (2004) New insights into vascular collapse and growth dynamics in solid tumors. J
Theor Biol 228(3):335–346
Association AD et al (2004) Screening for type 2 diabetes. Diabetes Care 27(suppl 1):s11–s14
Baker P, Mottram R (1973) Metabolism of exercising and resting human skeletal muscle, in the post-prandial
and fasting states. Clin Sci 44(5):479–491
Bielas J, Loeb K, Rubin B, True L, Loeb L (2006) Human cancers express a mutator phenotype. Proc Natl
Acad Sci 103(48):18238–18242
Boris J, Book D (1973) Flux-corrected transport. I. SHASTA, a fluid transport algorithm that works. J
Comput Phys 11(1):38–69
123
M. Kuznetsov, A. Kolobov
Boucher Y, Baxter L, Jain R (1990) Interstitial pressure gradients in tissue-isolated and subcutaneous tumors:
implications for therapy. Cancer Res 50(15):4478–4484
Bouchnita A, Belmaati F, Aboulaich R, Koury M, Volpert V (2017) A hybrid computation model to describe
the progression of multiple myeloma and its intra-clonal heterogeneity. Computation 5(1):16
Byrne H, Drasdo D (2009) Individual-based and continuum models of growing cell populations: a compar-
ison. J Math Biol 58(4–5):657
Casciari J, Sotirchos S, Sutherland R (1992) Mathematical modelling of microenvironment and growth in
EMT6/Ro multicellular tumour spheroids. Cell Prolif 25(1):1–22
Chisholm R, Lorenzi T, Lorz A, Larsen A, Almeida L, Escargueil A, Clairambault J (2015) Emergence of
drug tolerance in cancer cell populations: an evolutionary outcome of selection, non-genetic instability
and stress-induced adaptation. Cancer Res 75(6):930–939
Chmielecki J, Foo J, Oxnard G, Hutchinson K, Ohashi K, Somwar R, Wang L, Amato K, Arcila M, Sos M
et al (2011) Optimization of dosing for EGFR-mutant non-small cell lung cancer with evolutionary
cancer modeling. Sci Transl Med 3(90):90ra59–90ra59
Citron M, Berry D, Cirrincione C, Hudis C, Winer E, Gradishar W, Davidson N, Martino S, Livingston R,
Ingle J et al (2003) Randomized trial of dose-dense versus conventionally scheduled and sequential
versus concurrent combination chemotherapy as postoperative adjuvant treatment of node-positive
primary breast cancer: first report of intergroup Trial C9741/Cancer and Leukemia Group B Trial
9741. J Clin Oncol 21(8):1431–1439
Esteller M (2008) Epigenetics in cancer. New Engl J Med 358(11):1148–1159
Eymontt M, Gwinup G, Kruger F, Maynard D, Hamwi G (1965) Cushing’syndrome with hypoglycemia
caused by adrenocortical carcinoma. J Clin Endocrinol Metab 25(1):46–52
Fisher R (1937) The wave of advance of advantageous genes. Ann Eugen 7(4):355–369
Freyer J, Sutherland R (1985) A reduction in the in situ rates of oxygen and glucose consumption of cells
in EMT6/Ro spheroids during growth. J Cell Physiol 124(3):516–524
Gatenby R, Gawlinski E, Gmitro A, Kaylor B, Gillies R (2006) Acid-mediated tumor invasion: a multidis-
ciplinary study. Cancer Res 66(10):5216–5223
Gerlee P, Anderson A (2008) A hybrid cellular automaton model of clonal evolution in cancer: the emergence
of the glycolytic phenotype. J Theor Biol 250(4):705–722
Giese A, Loo M, Tran N, Haskett D, Coons S, Berens M (1996) Dichotomy of astrocytoma migration and
proliferation. Int J Cancer 67(2):275–282
Gottesman M (2002) Mechanisms of cancer drug resistance. Annu Rev Med 53(1):615–627
Greaves M, Maley C (2012) Clonal evolution in cancer. Nature 481(7381):306
Hadjiandreou M, Mitsis G (2014) Mathematical modeling of tumor growth, drug-resistance, toxicity, and
optimal therapy design. IEEE T Bio-Med Eng 61(2):415–425
Hanahan D, Weinberg R (2000) The hallmarks of cancer. Cell 100(1):57–70
Hanahan D, Weinberg R (2011) Hallmarks of cancer: the next generation. Cell 144(5):646–674
Hart D, Shochat E, Agur Z (1998) The growth law of primary breast cancer as inferred from mammography
screening trials data. Br J Cancer 78(3):382
Holash J, Maisonpierre P, Compton D, Boland P, Alexander C, Zagzag D, Yancopoulos G, Wiegand S
(1999) Vessel cooption, regression, and growth in tumors mediated by angiopoietins and VEGF.
Science 284(5422):1994–1998
Iwasa Y, Nowak M, Michor F (2006) Evolution of resistance during clonal expansion. Genetics 172(4):2557–
2566
Izuishi K, Kato K, Ogura T, Kinoshita T, Esumi H (2000) Remarkable tolerance of tumor cells to nutrient
deprivation: possible new biochemical target for cancer therapy. Cancer Res 60(21):6201–6207
Jiao Y, Torquato S (2011) Emergent behaviors from a cellular automaton model for invasive tumor growth
in heterogeneous microenvironments. PLoS Comput Biol 7(12):e1002314
Kathagen-Buhmann A, Schulte A, Weller J, Holz M, Herold-Mende C, Glass R, Lamszus K (2016) Glycol-
ysis and the pentose phosphate pathway are differentially associated with the dichotomous regulation
of glioblastoma cell migration versus proliferation. Neuro Oncol 18(9):1219–1229
Kolobov A, Kuznetsov M (2015) Investigation of the effects of angiogenesis on tumor growth using a
mathematical model. Biophysics 60(3):449–456
Kolobov A, Polezhaev A, Solyanik G (2000) The role of cell motility in metastatic cell dominance phe-
nomenon: analysis by a mathematical model. Comput Math Methods Med 3(1):63–77
Kuznetsov M, Kolobov A (2017) Mathematical modelling of chemotherapy combined with bevacizumab.
Russ J Numer Anal Model 32(5):293–304
123
Investigation of solid tumor progression with account...
Kuznetsov M, Kolobov A (2018) Transient alleviation of tumor hypoxia during first days of antiangiogenic
therapy as a result of therapy-induced alterations in nutrient supply and tumor metabolism-analysis
by mathematical modeling. J Theor Biol 451:86–100
Kuznetsov M, Gorodnova N, Simakov S, Kolobov A (2016) Multiscale modeling of angiogenic tumor
growth, progression, and therapy. Biophysics 61(6):1042–1051
Kuznetsov M, Gubernov V, Kolobov A (2018) Analysis of anticancer efficiency of combined fractionated
radiotherapy and antiangiogenic therapy via mathematical modelling. Russ J Numer Anal Model
33(4):225–242
Ledzewicz U, Naghnaeian M, Schättler H (2012) Optimal response to chemotherapy for a mathematical
model of tumor-immune dynamics. J Math Biol 64(3):557–577
Levick JR (2013) An introduction to cardiovascular physiology. Butterworth-Heinemann, Oxford
Lorenzi T, Chisholm R, Clairambault J (2016) Tracking the evolution of cancer cell populations through
the mathematical lens of phenotype-structured equations. Biol Direct 11(1):43
Lorenzi T, Venkataraman C, Lorz A, Chaplain M (2018) The role of spatial variations of abiotic factors in
mediating intratumour phenotypic heterogeneity. J Theor Biol 451:101–110
Lorz A, Lorenzi T, Hochberg M, Clairambault J, Perthame B (2013) Populational adaptive evolution,
chemotherapeutic resistance and multiple anti-cancer therapies. ESAIM Math Model Numer Anal
47(2):377–399
Lorz A, Lorenzi T, Clairambault J, Escargueil A, Perthame B (2015) Modeling the effects of space structure
and combination therapies on phenotypic heterogeneity and drug resistance in solid tumors. Bull Math
Biol 77(1):1–22
Louis D, Ohgaki H, Wiestler O, Cavenee W, Burger P, Jouvet A, Scheithauer B, Kleihues P (2007) The
2007 who classification of tumours of the central nervous system. Acta Neuropathol 114(2):97–109
Macklin P, McDougall S, Anderson A, Chaplain M, Cristini V, Lowengrub J (2009) Multiscale modelling
and nonlinear simulation of vascular tumour growth. J Math Biol 58(4–5):765–798
Moreno-Sánchez R, Rodríguez-Enríquez S, Marín-Hernández A, Saavedra E (2007) Energy metabolism in
tumor cells. FEBS J 274(6):1393–1418
Owen M, Alarcón T, Maini P, Byrne H (2009) Angiogenesis and vascular remodelling in normal and
cancerous tissues. J Math Biol 58(4–5):689
Patra K, Hay N (2014) The pentose phosphate pathway and cancer. Trends Biochem Sci 39(8):347–354
Phan L, Yeung S, Lee M (2014) Cancer metabolic reprogramming: importance, main features, and potentials
for precise targeted anti-cancer therapies. Cancer Biol Med 11(1):1
Press WH (2007) Numerical recipes. The art of scientific computing, 3rd edn. Cambridge University Press,
Cambridge
Robertson K (2001) Dna methylation, methyltransferases, and cancer. Oncogene 20(24):3139
Rockne R, Alvord E, Rockhill J, Swanson K (2009) A mathematical model for brain tumor response to
radiation therapy. J Math Biol 58(4–5):561
Shiraishi T, Verdone J, Huang J, Kahlert U, Hernandez J, Torga G, Zarif J, Epstein T, Gatenby R, McCart-
ney A et al (2015) Glycolysis is the primary bioenergetic pathway for cell motility and cytoskeletal
remodeling in human prostate and breast cancer cells. Oncotarget 6(1):130
Skehan P (1986) On the normality of growth dynamics of neoplasms in vivo: a data base analysis. Growth
50(4):496–515
Sonveaux P, Végran F, Schroeder T, Wergin M, Verrax J, Rabbani Z, De Saedeleer C, Kennedy K, Diepart
C, Jordan B et al (2008) Targeting lactate-fueled respiration selectively kills hypoxic tumor cells in
mice. J Clin Investig 118(12):3930
Stamatelos S, Kim E, Pathak A, Popel A (2014) A bioimage informatics based reconstruction of breast
tumor microvasculature with computational blood flow predictions. Microvasc Res 91:8–21
Stiehl T, Lutz C, Marciniak-Czochra A (2016) Emergence of heterogeneity in acute leukemias. Biol Direct
11(1):51
Strong L (1958) Genetic concept for the origin of cancer: historical review. Ann NY Acad Sci 71(6):810–838
Sudhakar A (2009) History of cancer, ancient and modern treatment methods. J Cancer Sci Ther 1(2):1
Swanson K, Alvord E Jr, Murray J (2000) A quantitative model for differential motility of gliomas in grey
and white matter. Cell Prolif 33(5):317–329
Theodorescu D, Cornil I, Sheehan C, Man S, Kerbel R (1991) Dominance of metastatically competent
cells in primary murine breast neoplasms is necessary for distant metastatic spread. Int J Cancer
47(1):118–123
123
M. Kuznetsov, A. Kolobov
Tuchin V, Bashkatov A, Genina E, Sinichkin Y, Lakodina N (2001) In vivo investigation of the immersion-
liquid-induced human skin clearing dynamics. Tech Phys Lett 27(6):489–490
Vander Heiden M, Cantley L, Thompson C (2009) Understanding the warburg effect: the metabolic require-
ments of cell proliferation. Science 324(5930):1029–1033
Velicescu M, Weisenberger D, Gonzales F, Tsai Y, Nguyen C, Jones P (2002) Cell division is required for
de novo methylation of CpG islands in bladder cancer cells. Cancer Res 62(8):2378–2384
Publisher’s Note Springer Nature remains neutral with regard to jurisdictional claims in published maps
and institutional affiliations.
Affiliations
Maxim Kuznetsov1·Andrey Kolobov1,2
1Division of Theoretical Physics, P.N. Lebedev Physical Institute of the Russian Academy of
Sciences, 53 Leninskii Prospekt, Moscow, Russia 119991
2Peoples’ Friendship University of Russia (RUDN University), 6 Miklukho-Maklaya St, Moscow,
Russia 117198
123
... Let us note that the models that utilize such approach generally neglect additional diffusion in phenotypic space. Spatially-distributed models of this type can however account for diffusion in physical space [43,68]. Some models that consider genetic instability use additional terms, nonlocal in phenotypic space, for description of competition between tumor cell populations [73,43,42]. ...
... However, their competition can arise from the model formulation naturally. An example, leading to such effect, is introduction of dependence of cell death rate on nutrient level in spatially-distributed case, which was implemented in the work [68]. Under such model formulation, the cells that cannot escape nutrient-deficient regions, naturally turn out to be less fit. ...
Article
Full-text available
Nonlocal reaction-diffusion equations describe various biological and biomedical applications. Their mathematical properties are essentially different in comparison with the local equations, and this difference can lead to important biological implications. This review will present the state of the art in the investigation of nonlocal reaction-diffusion models in biomedical applications. We will consider various models arising in mathematical immunology, neuroscience, cancer modelling, and we will discuss their mathematical properties, nonlinear dynamics, resulting spatiotemporal patterns and biological significance.
... Therefore, the global stability of the positive steady state as well as the tumor-free steady state cannot be strictly proved, which are left as open questions here. Since the model has many limitations due to our interests, a lot of modification can be consider, such as the conversion between proliferating and quiescent cells [24; 26], the diffusion of nutrition [35; 46], the movement of tumor cells [22], which leads more challenging in mathematical analysis. These may be investigated in the future work. ...
Preprint
Full-text available
A nonlinear age-structured tumor population model, in which the tumor population is divided into proliferating and quiescent cells, and the probability of mitosis in proliferating cells are determined by the concentration of nutrient and the proliferating cells, is provided and analyzed. The dynamic behaviors of the model are investigated by means of semigroup operator theory. Firstly, a threshold called $R_{0}$ is proposed, and the model always exists a tumor-free steady state, which is locally asymptotically stable if $R_{0}<1$, and a sufficient condition for its global stability is also obtained. Secondly, a special piecewise function with time delay describing the 'birth' of proliferating cells is introduced, and a particular threshold $R_{0s}$ in accordance with it is obtained. If $R_{0s}>1$ it is shown that there exists a unique positive steady state, which is locally asymptotically stable once it appears. Further, several numerical analyses are carried out to illustrate the validity of the theoretical results, which shows that the nonlinearity of split rate not only halts tumor's exponential growth but also forces the tumor population converge to constant and sometimes even like the Gompertz growth. The results suggest that it may be a good treatment to prolong the split time as long as possible during tumor therapy. Mathematics Subject Classification (2020) 34D20 · 35A25 · 92C37
... As cell migration and proliferation share common elements of molecular pathways, there might be competition in determining which process prevails in a given cell. Indeed, the mathematical modeling predicts that the increase in cell proliferation occurs at the expense of cell migration and vice versa [73]. This popular concept of the dichotomy between proliferation and migration is discussed in the literature as the "Go-or-Grow" or "Divide or conquer" concept. ...
Article
Full-text available
Epithelial mesenchymal transition (EMT) and mesenchymal epithelial transition (MET) are genetic determinants of cellular plasticity. These programs operate in physiological (embryonic development, wound healing) and pathological (organ fibrosis, cancer) conditions. In cancer, EMT and MET interfere with various signalling pathways at different levels. This results in gross alterations in the gene expression programs, which affect most, if not all hallmarks of cancer, such as response to proliferative and death-inducing signals, tumorigenicity, and cell stemness. EMT in cancer cells involves large scale reorganisation of the cytoskeleton, loss of epithelial integrity, and gain of mesenchymal traits, such as mesenchymal type of cell migration. In this regard, EMT/MET plasticity is highly relevant to the Go-or-Grow concept, which postulates the dichotomous relationship between cell motility and proliferation. The Go-or-Grow decisions are critically important in the processes in which EMT/MET plasticity takes the central stage, mobilisation of stem cells during wound healing, cancer relapse, and metastasis. Here we outline the maintenance of quiescence in stem cell and metastatic niches, focusing on the implication of EMT/MET regulatory networks in Go-or-Grow switches. In particular, we discuss the analogy between cells residing in hybrid quasi-mesenchymal states and GAlert, an intermediate phase allowing quiescent stem cells to enter the cell cycle rapidly.
... Its greatest value led to the maximum absolute value of adhesion stress close to σ cr , whereas at its minimum value, direct cell contact produced repulsion stress less than σ cr , which stimulated cell overlapping during active proliferation. The value for the tumor cell pair divergence speed v was obtained as r 0 /[ln 2/B], where B = 0.03 was a reasonable value for the tumor cell proliferation rate, which was also used in our previous works (see [43,44]). A single tumor cell division unconstrained by external forces happened in about 23 h. ...
Article
Full-text available
An off-lattice agent-based model of tumor growth is presented, which describes a tumor as a network of proliferating cells, whose dynamics depend on the stress generated by intercellular bonds. A numerical method is introduced that ensures the smooth dynamics of the cell network and allows for relative numerical cheapness while reproducing the effects typical of more complex approaches such as the elongation of cells toward low-pressure regions and their tendency to maximize the contact area. Simulations of free tumor growth, restricted only by the stress generated within the tumor, demonstrate the influence of the tissue hydraulic conductivity and strength of cell–cell interactions on tumor shape and growth rate. Simulations of compact tumor growth within normal tissue show that strong interaction between tumor cells is a major factor limiting tumor growth. Moreover, the effects of normal tissue size and strength of normal cell interactions on tumor growth are ambiguous and depend on the value of tissue hydraulic conductivity. Simulations of tumor growth in normal tissue with the account of nutrients yield different growth regimes, including growth without saturation for at least several years with the formation of large necrotic cores in cases of low tissue hydraulic conductivity and sufficiently high nutrient supply, which qualitatively correlates with known clinical data.
... The mathematical model of tumor growth and radiotherapy, considered in this work, represents an augmentation of the our previous model Kuznetsov and Kolobov (2020b), which in its turn is based on preceding versions, used for investigation of various aspects of tumor growth and treatment (Kuznetsov et al. (2018); Kuznetsov and Kolobov (2020a); Kuznetsov (2020Kuznetsov ( , 2021). The set of Eqs. ...
Article
Full-text available
A spatially-distributed continuous mathematical model of solid tumor growth and treatment by fractionated radiotherapy is presented. The model explicitly accounts for the factors, widely referred to as 4 R’s of radiobiology, which influence the efficacy of radiotherapy fractionation protocols: tumor cell repopulation, their redistribution in cell cycle, reoxygenation and repair of sublethal damage of both tumor and normal tissues. With the use of special algorithm the fractionation protocols that provide increased tumor control probability, compared to standard clinical protocol, are found for various physiologically-based values of model parameters under the constraints of fixed overall normal tissue damage and maximum admissible fractional dose. In particular, it is shown that significant gain in treatment efficacy can be achieved for tumors of low malignancy by the use of protracted hyperfractionated protocols. The optimized non-uniform protocols are characterized by gradual escalation of fractional doses in their last parts, which start after the levels of oxygen and nutrients significantly elevate throughout the tumor and accelerated tumor proliferation manifests itself, which is a well-known experimental phenomenon.
... The examples are the simulation of solid tumor progression towards an increasingly invasive phenotype [98] and the investigation of the effect of internalization of less motile cells into the tumor spheroid [99]. ...
Article
Full-text available
Despite significant advances in oncological research, cancer nowadays remains one of the main causes of mortality and morbidity worldwide. New treatment techniques, as a rule, have limited efficacy, target only a narrow range of oncological diseases, and have limited availability to the general public due their high cost. An important goal in oncology is thus the modification of the types of antitumor therapy and their combinations, that are already introduced into clinical practice, with the goal of increasing the overall treatment efficacy. One option to achieve this goal is optimization of the schedules of drugs administration or performing other medical actions. Several factors complicate such tasks: the adverse effects of treatments on healthy cell populations, which must be kept tolerable; the emergence of drug resistance due to the intrinsic plasticity of heterogeneous cancer cell populations; the interplay between different types of therapies administered simultaneously. Mathematical modeling, in which a tumor and its microenvironment are considered as a single complex system, can address this complexity and can indicate potentially effective protocols, that would require experimental verification. In this review, we consider classical methods, current trends and future prospects in the field of mathematical modeling of tumor growth and treatment. In particular, methods of treatment optimization are discussed with several examples of specific problems related to different types of treatment.
... Such approximation has been used in the previous works of our research group [32][33][34][35]. The conditions, under which this approximation is applicable, will be discussed in Section 4.1. ...
Article
Full-text available
A continuous mathematical model of non-invasive avascular tumor growth in tissue is presented. The model considers tissue as a biphasic material, comprised of a solid matrix and interstitial fluid. The convective motion of tissue elements happens due to the gradients of stress, which change as a result of tumor cells proliferation and death. The model accounts for glucose as the crucial nutrient, supplied from the normal tissue, and can reproduce both diffusion-limited and stress-limited tumor growth. Approximate tumor growth curves are obtained semi-analytically in the limit of infinite tissue hydraulic conductivity, which implies instantaneous equalization of arising stress gradients. These growth curves correspond well to the numerical solutions and represent classical sigmoidal curves with a short initial exponential phase, subsequent almost linear growth phase and a phase with growth deceleration, in which tumor tends to reach its maximum volume. The influence of two model parameters on tumor growth curves is investigated: tissue hydraulic conductivity, which links the values of stress gradient and convective velocity of tissue phases, and tumor nutrient supply level, which corresponds to different permeability and surface area density of capillaries in the normal tissue that surrounds the tumor. In particular, it is demonstrated, that sufficiently low tissue hydraulic conductivity (intrinsic, e.g., to tumors arising from connective tissue) and sufficiently high nutrient supply can lead to formation of giant benign tumors, reaching tens of centimeters in diameter, which are indeed observed clinically.
... System engineers thrive to provide a proper description of the tumor growth, however, the inter-and intrapatient variability and the effects of different drugs are cumbersome to model. Several models were introduced over the years [4][5][6][7][8][9]. In particular, we are concerned with the use of a third-order model. ...
Article
Full-text available
One challenging aspect of therapy optimization and application of control algorithms in the field of tumor growth modeling is the limited number of measurable physiological signals—state variables—and the knowledge of model parameters. A possible solution to provide such information is the application of observer or state estimator. One of the most widely applied estimators for nonlinear problems is the extended Kalman filter (EKF). In this study, a moving horizon estimation (MHE)-based observer is developed and compared to an optimized EKF. The observers utilize a third-order tumor growth model. The performance of the observers is tested on measurements gathered from a laboratory mice trial using chemotherapeutic drug. The proposed MHE is designed to be suitable for closed-loop applications and yields simultaneous state and parameter estimation.
... The mathematical model of tumor growth, considered herein, was based on our models, previously used for the investigation of various aspects of tumor growth and treatment [20][21][22][23]. There were five variables in this version of the model, which were the functions of space and time coordinates, r and t: the density of tumor cells n(r, t), the density of normal cells h(r, t), the fraction of necrotic tissue m(r, t), the concentration of glucose g(r, t) and the concentration of oxygen ω(r, t). ...
Article
Full-text available
A spatially-distributed continuous mathematical model of solid tumor growth and treatment by fractionated radiotherapy is presented. The model explicitly accounts for three time and space-dependent factors that influence the efficiency of radiotherapy fractionation schemes—tumor cell repopulation, reoxygenation and redistribution of proliferative states. A special algorithm is developed, aimed at finding the fractionation schemes that provide increased tumor cure probability under the constraints of maximum normal tissue damage and maximum fractional dose. The optimization procedure is performed for varied radiosensitivity of tumor cells under the values of model parameters, corresponding to different degrees of tumor malignancy. The resulting optimized schemes consist of two stages. The first stages are aimed to increase the radiosensitivity of the tumor cells, remaining after their end, sparing the caused normal tissue damage. This allows to increase the doses during the second stages and thus take advantage of the obtained increased radiosensitivity. Such method leads to significant expansions in the curative ranges of the values of tumor radiosensitivity parameters. Overall, the results of this study represent the theoretical proof of concept that non-uniform radiotherapy fractionation schemes may be considerably more effective that uniform ones, due to the time and space-dependent effects.
Article
Oral squamous cell carcinoma (OSCC) has become the most common HPV-related cancer with high invasion and metastasis. Exploring biomarkers for the screening and monitoring of OSCC, especially for the HPV-OSCC, would benefit patients’ diagnosis and prognosis. This study evaluated the significance and mechanism of TMEM161B-AS1 and miR-651-5p in HPV-OSCC aiming to provide novel insight into the mechanism of HPV-OSCC development. Expression of TMEM161B-AS1 and miR-561-5p was analyzed in healthy individuals, HPV-infected non-OSCC patients, and HPV-OSCC patients using PCR. Their significance in HPV-OSCC occurrence and prognosis was evaluated by logistic regression, ROC, Kaplan–Meier, and Cox regression analysis. In OSCC cells, CCK8 and Transwell assays were employed for assessing cell growth and metastasis. The luciferase reporter assay and cell transfection were performed to evaluate the regulatory association between TMEM161B-AS1, miR-561-5p, and BDNF. Significant upregulation of TMEM161B-AS1 and downregulation of miR-561-5p were observed in oral HPV-infected patients. Both TMEM161B-AS1 and miR-651-5p served as risk factors for the occurrence of OSCC in oral HPV-infected patients and could distinguish HPV-OSCC patients from HPV-infected non-OSCC patients. Increased TMEM161B-AS1 and reduced miR-561-5p indicated severe development and adverse prognosis of HPV-OSCC patients. In OSCC cells, silencing TMEM161-AS1 suppressed cell proliferation and motility via negatively modulating miR-561-5p. miR-561-5p negatively regulated BDNF, which was considered the underlying mechanism of TMEM161B-AS1. Increasing TMEM161B-AS expression and decreasing miR-561-5p showed the occurrence of OSCC in HPV-infected patients and predicted malignant development and adverse prognosis. TMEME161B-AS1 served as a tumor promoter via regulating the miR-561-5p/BDNF axis.
Article
Full-text available
One of the relatively new modalities in treatment of cancer is antiangiogenic therapy (AAT), which stops the formation of new blood vessels, thus leading to nutrient deprivation of a tumor. It has been discovered that it can cause transient alleviation of intratumoral hypoxia, which has induced interest in investigation of its combination with radiotherapy (RT), since presence of oxygen stimulates efficiency of the latter. Preclinical and clinical studies have shown ambiguous results concerning such combined treatment, indicating a need for theoretical investigation. For that purpose, we have developed a spatially-distributed mathematical model of tumor growth and combined RT with AAT. Model simulations in a physiologically justified range of parameters suggest that outcome of addition of AAT to RT should depend on the tumor radiosensitivity: under its low values, when RT de facto serves as a palliative therapy, inclusion of AAT may prolong patient’s survival; however, under high tumor radiosensitivity, AAT may compromise curative effect of RT.
Article
Full-text available
Antiangiogenic therapy is aimed at the blocking of angiogenesis, i.e., the process of new blood vessels formation, which should decrease oxygen and nutrients inflow to tumor and thus slow down its growth. This type of therapy is frequently administered together with chemotherapy, which kills rapidly proliferating tumor cells, as well as other dividing cells of the body, thus leading to significant adverse effects. However, action of bevacizumab inevitably influences the inflow of chemotherapeutic drug in tumor, therefore, optimal scheduling of drug administration is an important problem. Using a model based on consideration of the most crucial processes happening during tumor progression and therapy, we compare effectiveness of different schemes of combined chemotherapy with bevacizumab and propose new scheme of drug administration which may lead to enhanced antitumor effect compared to classical clinical scheme of simultaneous drug administration.
Article
Full-text available
The chronic Philadelphia-negative myeloproliferative neoplasms (MPNs) are acquired stem cell neoplasms which ultimately may transform to acute myelogenous leukemia. Most recently, chronic inflammation has been described as an important factor for the development and progression of MPNs in the biological continuum from early cancer stage to the advanced myelofibrosis stage, the MPNs being described as “A Human Inflammation Model for Cancer Development“. This novel concept has been built upon clinical, experimental, genomic, immunological and not least epidemiological studies. Only a few studies have described the development of MPNs by mathematical models, and none have addressed the role of inflammation for clonal evolution and disease progression. Herein, we aim at using mathematical modelling to substantiate the concept of chronic inflammation as an important trigger and driver of MPNs.The basics of the model describe the proliferation from stem cells to mature cells including mutations of healthy stem cells to become malignant stem cells. We include a simple inflammatory coupling coping with cell death and affecting the basic model beneath. First, we describe the system without feedbacks or regulatory interactions. Next, we introduce inflammatory feedback into the system. Finally, we include other feedbacks and regulatory interactions forming the inflammatory-MPN model. Using mathematical modeling, we add further proof to the concept that chronic inflammation may be both a trigger of clonal evolution and an important driving force for MPN disease progression. Our findings support intervention at the earliest stage of cancer development to target the malignant clone and dampen concomitant inflammation.
Article
Full-text available
A mathematical model of angiogenic tumor growth in tissue with account of bevacizumab therapy was developed. The model accounts for convective flows that occur in dense tissue under active division of tumor cells, as well as the migration and proliferation dichotomy of malignant cells, which depends on the concentrations of major metabolites, such as oxygen and glucose. Tumor cells wich are in a state of metabolic stress produce vascular endothelial growth factor, which stimulates angiogenesis. To establish the relationship between the capillary network density and oxygen supply, a separate model of stationary blood flow in the capillary network was developed and investigated. A numerical study of the tumor-growth model showed that antiangiogenic bevacizumab treatment of tumors of the diffuse type reduces the total number of their cells, but practically does not affect the rate of their invasion into normal tissues. At the same time, it was found that the growth of dense tumors may be non-monotonic in a rather wide range of parameters. It was shown that in this case bevacizumab therapy stabilizes and significantly inhibits tumor growth, while its local-in-time efficiency is sensitive to the time that it begins.
Article
Full-text available
Background Leukemias are malignant proliferative disorders of the blood forming system. Sequencing studies demonstrate that the leukemic cell population consists of multiple clones. The genetic relationship between the different clones, referred to as the clonal hierarchy, shows high interindividual variability. So far, the source of this heterogeneity and its clinical relevance remain unknown. We propose a mathematical model to study the emergence and evolution of clonal heterogeneity in acute leukemias. The model allows linking properties of leukemic clones in terms of self-renewal and proliferation rates to the structure of the clonal hierarchy. Results Computer simulations imply that the self-renewal potential of the first emerging leukemic clone has a major impact on the total number of leukemic clones and on the structure of their hierarchy. With increasing depth of the clonal hierarchy the self-renewal of leukemic clones increases, whereas the proliferation rates do not change significantly. The emergence of deep clonal hierarchies is a complex process that is facilitated by a cooperativity of different mutations. Conclusion Comparison of patient data and simulation results suggests that the self-renewal of leukemic clones increases with the emergence of clonal heterogeneity. The structure of the clonal hierarchy may serve as a marker for patient prognosis. Reviewers This article was reviewed by Marek Kimmel, Tommaso Lorenzi and Tomasz Lipniacki.
Article
We present here a space- and phenotype-structured model of selection dynamics between cancer cells within a solid tumour. In the framework of this model, we combine formal analyses with numerical simulations to investigate in silico the role played by the spatial distribution of abiotic components of the tumour microenvironment in mediating phenotypic selection of cancer cells. Numerical simulations are performed both on the 3D geometry of an in silico multicellular tumour spheroid and on the 3D geometry of an in vivo human hepatic tumour, which was imaged using computerised tomography. The results obtained show that inhomogeneities in the spatial distribution of oxygen, currently observed in solid tumours, can promote the creation of distinct local niches and lead to the selection of different phenotypic variants within the same tumour. This process fosters the emergence of stable phenotypic heterogeneity and supports the presence of hypoxic cells resistant to cytotoxic therapy prior to treatment. Our theoretical results demonstrate the importance of integrating spatial data with ecological principles when evaluating the therapeutic response of solid tumours to cytotoxic therapy.
Article
A number of experiments on mouse tumor models, as well as certain clinical data, have demonstrated, that antiangiogenic therapy can lead to transient improvement in tumor oxygenation, that allows to increase efficiency of following radiotherapy. In the majority of works, this phenomenon has been explained by enhanced tumor perfusion due to normalization of capillaries' structure, that results in elevated oxygen inflow in tumor. However, changes in tumor perfusion often haven't been directly measured in relevant works, moreover, antiangiogenic therapy has been proven to have ambiguous effect on tumor perfusion both in mouse tumor models and in clinics. Herein, we suggest that elevation of blood perfusion may be not the only reason for transient alleviation of tumor hypoxia, and that it may manifest itself even under unchanged tumor blood flow. We propose that it may be as well caused by the decrease in tumor oxygen consumption rate (OCR) due to the reduction of tumor proliferation level, caused by nutrient shortage in result of antiangiogenic treatment. We provide detailed explanation of this hypothesis and visualize it using a specially developed mathematical model, which takes into account basic features of tumor growth and antiangiogenic therapy. We investigate the influence of the model parameters on oxygen dynamics; demonstrate, that transient alleviation of tumor hypoxia occurs in a fairly wide range of physiologically justified values of parameters; and point out the major factors, that determine oxygen dynamics during antiangiogenic therapy.
Article
Multiplemyeloma(MM)isageneticallycomplexhematologicalcancerthatischaracterized by proliferation of malignant plasma cells in the bone marrow. MM evolves from the clonal premalignant disorder monoclonal gammopathy of unknown significance (MGUS) by sequential genetic changes involving many different genes, resulting in dysregulated growth of multiple clones of plasma cells. The migration, survival, and proliferation of these clones require the direct and indirect interactions with the non-hematopoietic cells of the bone marrow. We develop a hybrid discrete-continuous model of MM development from the MGUS stage. The discrete aspect of the modelisobservedatthecellularlevel: cellsarerepresentedasindividualobjectswhichmove,interact, divide, and die by apoptosis. Each of these actions is regulated by intracellular and extracellular processes as described by continuous models. The hybrid model consists of the following submodels that have been simplified from the much more complex state of evolving MM: cell motion due to chemotaxis, intracellular regulation of plasma cells, extracellular regulation in the bone marrow, and acquisition of mutations upon cell division. By extending a previous, simpler model in which the extracellular matrix was considered to be uniformly distributed, the new hybrid model provides a more accurate description in which cytokines are produced by the marrow microenvironment and consumed by the myeloma cells. The complex multiple genetic changes in MM cells and the numerous cell-cell and cytokine-mediated interactions between myeloma cells and their marrow microenviroment are simplified in the model such that four related but evolving MM clones can be studied as they compete for dominance in the setting of intraclonal heterogeneity.