Processing math: 100%
Research article

Environmental stress level to model tumor cell growth and survival

  • Received: 17 January 2022 Revised: 14 March 2022 Accepted: 16 March 2022 Published: 28 March 2022
  • Survival of living tumor cells underlies many influences such as nutrient saturation, oxygen level, drug concentrations or mechanical forces. Data-supported mathematical modeling can be a powerful tool to get a better understanding of cell behavior in different settings. However, under consideration of numerous environmental factors mathematical modeling can get challenging. We present an approach to model the separate influences of each environmental quantity on the cells in a collective manner by introducing the "environmental stress level". It is an immeasurable auxiliary variable, which quantifies to what extent viable cells would get in a stressed state, if exposed to certain conditions. A high stress level can inhibit cell growth, promote cell death and influence cell movement. As a proof of concept, we compare two systems of ordinary differential equations, which model tumor cell dynamics under various nutrient saturations respectively with and without considering an environmental stress level. Particle-based Bayesian inversion methods are used to quantify uncertainties and calibrate unknown model parameters with time resolved measurements of in vitro populations of liver cancer cells. The calibration results of both models are compared and the quality of fit is quantified. While predictions of both models show good agreement with the data, there is indication that the model considering the stress level yields a better fitting. The proposed modeling approach offers a flexible and extendable framework for considering systems with additional environmental factors affecting the cell dynamics.

    Citation: Sabrina Schönfeld, Alican Ozkan, Laura Scarabosio, Marissa Nichole Rylander, Christina Kuttler. Environmental stress level to model tumor cell growth and survival[J]. Mathematical Biosciences and Engineering, 2022, 19(6): 5509-5545. doi: 10.3934/mbe.2022258

    Related Papers:

    [1] Rafael Martínez-Fonseca, Cruz Vargas-De-León, Ramón Reyes-Carreto, Flaviano Godínez-Jaimes . Bayesian analysis of the effect of exosomes in a mouse xenograft model of chronic myeloid leukemia. Mathematical Biosciences and Engineering, 2023, 20(11): 19504-19526. doi: 10.3934/mbe.2023864
    [2] Hsiu-Chuan Wei . Mathematical modeling of tumor growth: the MCF-7 breast cancer cell line. Mathematical Biosciences and Engineering, 2019, 16(6): 6512-6535. doi: 10.3934/mbe.2019325
    [3] Elena Izquierdo-Kulich, Margarita Amigó de Quesada, Carlos Manuel Pérez-Amor, Magda Lopes Texeira, José Manuel Nieto-Villar . The dynamics of tumor growth and cells pattern morphology. Mathematical Biosciences and Engineering, 2009, 6(3): 547-559. doi: 10.3934/mbe.2009.6.547
    [4] Filippo Cacace, Valerio Cusimano, Alfredo Germani, Pasquale Palumbo, Federico Papa . Closed-loop control of tumor growth by means of anti-angiogenic administration. Mathematical Biosciences and Engineering, 2018, 15(4): 827-839. doi: 10.3934/mbe.2018037
    [5] Avner Friedman, Yangjin Kim . Tumor cells proliferation and migration under the influence of their microenvironment. Mathematical Biosciences and Engineering, 2011, 8(2): 371-383. doi: 10.3934/mbe.2011.8.371
    [6] Yuyang Xiao, Juan Shen, Xiufen Zou . Mathematical modeling and dynamical analysis of anti-tumor drug dose-response. Mathematical Biosciences and Engineering, 2022, 19(4): 4120-4144. doi: 10.3934/mbe.2022190
    [7] Samantha L Elliott, Emek Kose, Allison L Lewis, Anna E Steinfeld, Elizabeth A Zollinger . Modeling the stem cell hypothesis: Investigating the effects of cancer stem cells and TGF−β on tumor growth. Mathematical Biosciences and Engineering, 2019, 16(6): 7177-7194. doi: 10.3934/mbe.2019360
    [8] Elena Izquierdo-Kulich, José Manuel Nieto-Villar . Morphogenesis of the tumor patterns. Mathematical Biosciences and Engineering, 2008, 5(2): 299-313. doi: 10.3934/mbe.2008.5.299
    [9] Mohammad A. Tabatabai, Wayne M. Eby, Karan P. Singh, Sejong Bae . T model of growth and its application in systems of tumor-immunedynamics. Mathematical Biosciences and Engineering, 2013, 10(3): 925-938. doi: 10.3934/mbe.2013.10.925
    [10] Maria Vittoria Barbarossa, Christina Kuttler, Jonathan Zinsl . Delay equations modeling the effects of phase-specific drugs and immunotherapy on proliferating tumor cells. Mathematical Biosciences and Engineering, 2012, 9(2): 241-257. doi: 10.3934/mbe.2012.9.241
  • Survival of living tumor cells underlies many influences such as nutrient saturation, oxygen level, drug concentrations or mechanical forces. Data-supported mathematical modeling can be a powerful tool to get a better understanding of cell behavior in different settings. However, under consideration of numerous environmental factors mathematical modeling can get challenging. We present an approach to model the separate influences of each environmental quantity on the cells in a collective manner by introducing the "environmental stress level". It is an immeasurable auxiliary variable, which quantifies to what extent viable cells would get in a stressed state, if exposed to certain conditions. A high stress level can inhibit cell growth, promote cell death and influence cell movement. As a proof of concept, we compare two systems of ordinary differential equations, which model tumor cell dynamics under various nutrient saturations respectively with and without considering an environmental stress level. Particle-based Bayesian inversion methods are used to quantify uncertainties and calibrate unknown model parameters with time resolved measurements of in vitro populations of liver cancer cells. The calibration results of both models are compared and the quality of fit is quantified. While predictions of both models show good agreement with the data, there is indication that the model considering the stress level yields a better fitting. The proposed modeling approach offers a flexible and extendable framework for considering systems with additional environmental factors affecting the cell dynamics.



    As cancerous diseases still cause many deaths in human population, there is a high need of improving the understanding of tumor growth dynamics and treatment strategies. For that purpose and for treatment optimization, mathematical approaches can help with adequate quantitative descriptions and predictions. Numerous mathematical models and methods are already available (see [1,2]), including differential equations of all types, stochastic models, phenomenological models as well as more mechanistic ones. Despite good new experimental techniques, e.g. measuring tumor growth also in vivo [3,4,5], it is still difficult to quantify growth by determining parameter values in realistic situations without harming a patient or affecting a potential treatment. To get a better quantitative understanding of the underlying processes, investigating tumor cell cultures and even single cells under well-controlled conditions provides a promising way to get at least parts of the necessary information (see e.g. [6]). But even under laboratory conditions, parameters can underlie variations and cells may behave individually to some extent. To consider this, uncertainty quantification can be an adequate tool. Having such in vitro experiments, it is much easier to separate and measure different environmental influences on the tumor growth. The probably most relevant environmental influences are the availability of nutrients and oxygen. For modeling purposes, one often focuses on the limiting nutrient, without specifying it in detail. Nutrient deprivation causes "stress" to any cells, but other factors may cause stress to the cells, as well. Thus, it is worth to consider the concept of stress as a relevant factor for modeling growth and death of tumor cells.

    To approach this problem, we suggest the introduction of an "environmental stress level" as an immeasurable auxiliary variable, which quantifies to what extent viable cells would get stressed out under certain environmental conditions. These may include quantities like nutrient saturation, oxygen level, drug concentrations or mechanical forces. On the one hand, this environmental stress level can e.g. inhibit cell growth and promote death by inducing necrosis for present cells. On the other hand, it is assumed to have an influence on cell movement, as cells favor areas with a low environmental stress level, which may cause metastasis formation, as well. In this work, this approach is presented in a simplified spatially homogeneous setting, where only the nutrient saturation effects the environmental stress level. Nevertheless, the presented model could be easily extended to more environmental variables, which provides several advantages from a modeling point of view. Even in this very simplified setting, model comparison shows a slight preference for the newly introduced model.

    For comparison, we present two systems of ordinary differential equations, which model tumor cell population dynamics respectively with and without considering an environmental stress level. To do so, time-resolved in vitro data (from [7]) are used to perform model parameter calibration with Bayesian inversion. The latter approach provides a quantitative estimate of the uncertainty in the estimated parameters and hence brings more information than a deterministic inversion method, at the price of a higher computational cost. In the Bayesian inversion setting several methods are available for sampling. Among the most popular ones are Markov Chain Monte Carlo (MCMC) methods [8, Ch. 6-7], which are very simple to implement, but they need ad hoc tuning when the dimension of the parameter space is high or for complex posterior distributions [9]. In the last decade, particle based methods have gained popularity for their higher efficiency and robustness compared to simple MCMC, especially in the presence of high dimensional parameter spaces and multimodal posterior distributions [10]. Most of these algorithms are population based, which means they deal with a collection of samples in every iteration. In this category are sequential importance sampling and resampling (annealed importance sampling [11] and Sequential Monte Carlo [12,13]) and population MCMC [14,15]. To calibrate the unknown parameters of the presented models, we use Sequential Monte Carlo (SMC). It is worth mentioning that the ensemble Kalman filter [16] can also be used to solve the inverse problem. It can be viewed as an approximate SMC with Gaussian approximation of the posterior [17]. However, the rigorous analysis of this method is confined for the moment to linear forward operators [17,18]. Eventually, the calibration results are used to compare the models via a validation metric and the Bayes factor. For this purpose, the chosen SMC method is advantageous opposed to other algorithms (e.g. MCMC), since its structure provides easy access to the Bayes factor [10].

    The article is structured as follows: The upcoming Section 2 explains the investigated mathematical models with the underlying experimental setup and provides background about the applied methods for uncertainty quantification, parameter calibration and model comparison. The results from the model calibrations are summarized and discussed in Section 3. Finally, Section 4 concludes the presented results and the use of the novel modeling approach. At the end of this article, more details about mathematical analysis of the models and the calibration results are enclosed in a supplementary section.

    Section 2.1 presents different models for population dynamics of tumor cells and introduces the idea of using a stress level to model environmental influences on the cells collectively. To investigate the effect of nutrient changes on viable cells, we consider a special case of the models. The underlying biological setting and the measurement methods for the corresponding in vitro experiments are explained in Section 2.2. With these data we calibrate the unknown parameters of the models using Bayesian inversion and Sequential Monte Carlo methods. Section 2.3 provides the theoretical background behind the algorithms. Finally, we present the validation metric and the Bayes factor in Section 2.4, which are used to investigate and compare the resulting model calibrations.

    The presented ODE models describe the basic growth and death dynamics of tumor cells in a spatially homogeneous, avascular environment under consideration of the present nutrient saturation. The time-dependent variables are the density of viable tumor cells V=V(t), the nutrient saturation S=S(t), and the environmental stress level η=η(t) at day t0.

    We consider two possibilities to model the influence of the nutrient saturation on the cells: Model MS assumes that the nutrient saturation directly affects the growth/death rates of the cells, while in model Mη the nutrient supply influences the cell population indirectly by changing the environmental stress level, which itself has an effect on the viability of the cells. The nutrient saturation is bounded and normalized: S(t)[0,1]. The bounds represent the complete absence of nutrients (S=0) and a nutrient supply, which creates optimal growth conditions for the cells (S=1). In the latter case, both models can be reduced to a simplified model, which is independent of the nutrient saturation S. We denote this model by Mopt ("opt" short for "optimal" growth conditions).

    A detailed description of each model follows in the consecutive paragraphs. All models are biologically reasonable in a first check, as they have been mathematically analyzed in terms of positivity and boundedness of the solutions as well as steady states and their stability. The last paragraph of this section provides a summary of the variables, parameters, and mathematical features of the models.

    What if the cells find optimal nutrient conditions? External nutrients serve as an energy source for cell proliferation. Growth can happen with a maximal possible proliferation rate β, which is virtually reached under optimal nutrient conditions (S=1). Under these circumstances, cell death is assumed to occur only for nutrient-independent reasons (e.g. dying of old age) with a constant rate λ. These dynamics can be modeled with a combination of a generalized logistic growth term (first proposed in [19]) and an exponential death term. The population growth is limited by the carrying capacity K of the biological system, taking into account e.g. limited space but not nutrient shortage. In particular, cell growth is inhibited in a cell density dependent manner, which is known as "proliferation contact inhibition" [20]. The strength of this phenomenon is specific for the cell type and is modeled by the shape parameter m of the logistic growth. A small value for m indicates strong contact inhibition, i.e. reduction of proliferation already starts at small cell densities. Therefore, 1/m can be interpreted as the strength of contact inhibition. Overall, this results in the following initial value problem:

    S=1{˙V=βV(1(VK)m)λV,V(0)=V0. (Mopt)

    As a feature of the pure logistic term, population growth for very small populations (i.e. VK) can be approximated by exponential growth with rate β. It is a reasonable assumption that, independently of the initial population size V0, the cell population's size actually increases over time under optimal nutrient conditions, which translates to a parameter relation: β>λ. With this constraint, the ODE can be rewritten as a purely logistic growth term with a new rate ˜β and capacity ˜K:

    ˙V=βV(1λβ(VK)m)β>λ=βV(1λβ)(1(VK(1λβ)1/m)m)=(βλ)=˜βV(VK(1λβ)1/m1(VK(1λβ)1/mVK(1λβ)1/m=˜KVK(1λβ)1/m)mVK(1λβ)1/m).

    Therefore, under optimal nutrient conditions, the cell population actually grows logistically with

    "net" growth rate:˜β=(βλ)(0,β)and "net" carrying capacity:˜K=K(1λβ)1/m=K(βλβ)1/m(0,K).

    For t0 the known analytical solution of such an initial value problem, i.e. for model Mopt, is

    V(t)=V0˜K(V0m+(˜KmV0m)e˜βmt)1/m=((V0K)m(βλ)βV0m+((βλ)KmβV0m)e(βλ)mt)1/m. (2.1)

    What if suboptimal nutrient conditions directly affect the cell dynamics? Exposing the cells to suboptimal nutrient conditions (0S<1) can affect the cells' metabolism by decelerating proliferation or even lead to cell death by starvation. To include this, an additional death term is needed, whose rate as well as the proliferation rate are scaled appropriately to the nutrient supply.

    A time-dependent nutrient supply can be described by adding an additional differential equation to the system: ˙S=g(S,V,t) with S(0)=S0. The reaction term g(S,V,t) can e.g. include terms for nutrient consumption by viable cells or an external nutrient source. In the considered in vitro experiments, whose measurements are used to calibrate the model later on, the nutrient concentration is maintained approximately constant. Therefore, all presented models assume

    ˙S=g(S,V,t)=0S(t)=const.=S0[0,1]t0.

    The present nutrient saturation S0 contributes to the cells' reproductivity as well as to their ability to survive. If there are no nutrients available, the cells die from starvation with a maximal possible rate λst, assumed to happen in an exponential manner. With increasing nutrient supply, the starvation rate λst decreases and the proliferation rate β is up-regulated. Adapting the previous model Mopt accordingly yields the more general initial value problem

    {˙V=δ+(S0)β=βSV(1(VK)m)(λ+δ(S0)λst)=λSV,V(0)=V0, (MS)

    with appropriate nutrient-dependent functions δ+(S) and δ(S), scaling the total growth and death rates:

    βS=δ+(S0)βandλS=(λ+δ(S0)λst).

    The scaling functions δ±(S) need to fulfill certain criteria to be biologically meaningful in our model setting. First, the cells should not die from starvation but proliferate with the maximal possible rate under optimal nutrient conditions. In contrast, without nutrients the cells should not be able to reproduce but starve with a maximal possible rate. These features translate to the properties

    δ+(S)β{βfor S1,(0,β)for 0<S<1,0for S0,andδ(S)λst{0for S1,(0,λst)for 0<S<1,λstfor S0.

    Second, increasing the nutrient saturation should promote growth and reduce starvation: δ+(S0) resp. δ(S0) need to be monotonically increasing resp. decreasing functions bounded by [0,1]. Therefore, a possible choice are Hill type functions with Hill coefficient k=2, i.e.

    δ+:SS2Sthr2+S2andδ:S1δ+(S)=1S2Sthr2+S2, (2.2)

    where 0<Sthr1 denotes the nutrient threshold, for which the cells proliferate/starve with half-maximal rate. For simplicity, we do not assume any hysteresis effect, motivating the relation δ(S)=1δ+(S). The scaling functions δ±(S) can be interpreted as influence functions, which describe how tolerant viable cells are to nutrient changes. In this context, the parameter Sthr can be seen as a nutrient sensitivity threshold. Figure 1 shows a qualitative plot of the behavior of the scaling functions.

    Figure 1.  Qualitative behavior of the Hill type scaling functions δ+(S) and δ(S).

    It can be shown (for calculations see Section 4.1 in the supplement) that for t0 a well-defined analytical solution of problem (MS) is

    V(t)={V0K(βSλSβSV0m+((βSλS)KmβSV0m)e(βSλS)mt)1/mif βSλS,V0K(1mtβSV0m+Km)1/mif βS=λS.

    We note that in the extreme situation of having no nutrient supply (i.e. S0=0), it holds δ+(S0)=0 and δ(S0)1. This results in βS=0 and λSλ+λst>0. By inserting this special case of βSλS into the given analytical solution, it simplifies to

    V(t)S00V0K((λ+λst)(λ+λst)Kme(λ+λst)mt)1/m=V0e(λ+λst)t.

    This is an expected result, being the solution of a simple exponential decay model without growth:

    {˙V=(λ+λst)V,V(0)=V0.

    Advantages of using the environmental stress level (ESL) instead to directly affect the cells. For an alternative way to include the effect of the nutrient conditions in model Mopt, we introduce an auxiliary variable: the environmental stress level η=η(t). It is designed to be an immeasurable quantity, describing to what extent viable cells would get stressed out, if exposed to certain environmental conditions, like in this case the nutrient saturation. We assume that the stress level is limited by a maximal level, which is set to one: η[0,1].

    In a general setting, the the ESL can be influenced by multiple environmental factors (e.g nutrient/oxygen/drug concentration), where each can be mathematically represented by a system variable. Let E1(t),,En(t) be a set of such environmental variables, with

    {˙E1=g1(E1,,En,V,t),E1(0)=E1,0,˙En=gn(E1,,En,V,t),En(0)=En,0,

    being the given reaction equations. Then the ODE for the environmental stress level can be given by

    ˙η=(nj=1αjδj(Ej))(1η) increasing stress level  (stressful conditions) (nj=1α+jδ+j(Ej))η recovery from stress  (beneficial conditions) with η(0)=η0[0,1]. (2.3)

    The first term includes the dynamics raising the ESL. Since η[0,1], the increase is bounded by one from above via the factor (1η). If the environmental factors generate beneficial growth conditions for the cells, the ESL decreases according to the second term. The parameters αj resp. α+j are variable-specific "sensitivity rates", denoting how fast the stress level increases/decreases, if the value of the associated variable Ej enters/leaves a range, which is critical for survival. With distinguishing between α and α+ per variable, it is possible to mathematically consider that cells react e.g. faster to the lack of an environmental factor, which is essential for their survival, than they recover from the deprivation once beneficial conditions are restored. Which values of Ej are considered to be critical, is given by the explicit shape of the corresponding "influence functions" δj(Ej) and δ+j(Ej). In particular, large values of δj(Ej) promote stress, whereas large values of δ+j(Ej) inhibits stress.

    Collecting the influences of the environment in the stress level gives rise to some practical advantages. We can choose an arbitrary number of environmental variables and include them easily into the stress reaction due to the modular structure of (2.3). With the parameters αj and α+j we can consider different sensitivities of the cells to changes in different environmental factors. Although neither these parameters nor the stress level η can be measured directly in experiments, appropriate data of cell viability can provide enough information to estimate the values of the sensitivity rates. For this e.g. statistical methods for parameter calibration can be used, where the only prior assumption is a rough range, in which the parameter values are expected. Then, by grouping these sensitivity rates according to their magnitude, we can distinguish the corresponding environmental variables between having a fast, slow or even no impact on the cells. Under a quasi-steady state assumption, the ODE for η can then be simplified by choosing a time scale of interest and omitting environmental variables with no influence.

    In the considered experimental setting, we have the special case of the constant nutrient saturation S(t) being the only environmental variable:

    E1(t)=S(t)=S0withg1(E1,V,t)=g(S,V,t)=0.

    We assume the ESL to increase and decrease with the same "nutrient sensitivity rate" αS. The corresponding "nutrient influence functions" are then δ(S) and δ+(S), as introduced previously in (2.2) for model MS. Hence, we have

    α1=α+1=αS,δ1(E1)=δ(S0)andδ+1(E1)=δ+(S0)=1δ(S0),

    i.e. given the initial condition η(0)=η0[0,1] we consider the ODE

    ˙η=αSδ(S0)(1η)αS(1δ(S0))η=αS(δ(S0)η).

    Since this equation is independent of the variable V, the exact solution can be calculated by separation of variables and for t0 it is given as:

    η(t)=δ(S0)(1eαSt)+η0eαSt. (2.4)

    Instead of the nutrient-dependent scaling functions δ± in the basic model MS, now the ESL influences the proliferation and starvation rate. A high stress level causes slower proliferation and faster starvation, whereas a low stress level has the opposite effect. Since by definition η(t)[0,1], this results in the following initial value problem:

    {˙V=(1η(t))β=βη(t)V(1(VK)m)(λ+η(t)λst)=λη(t)V,˙η=αSδ(S0)(1η)αS(1δ(S0))η,V(0)=V0,η(0)=η0. (Mη)

    By inserting the explicit time-dependent analytical solution (2.4), the system reduces to a non-autonomous ODE for V, which is solved numerically. We observe that model Mη has an interesting relation to the model MS: under the assumption that critical nutrient changes immediately influence the viable cells, i.e. assuming αS, we return to model MS, since for t0 it holds

    η(t)(2.4)=δ(S0)(1eαSt)+η0eαStαSδ(S0)=const.βη(t)αSβS,λη(t)αSλS.

    The same result can be achieved by a quasi-steady state assumption, stating that changes in the ESL happen at a much faster time scale than changes in the tumor cell number. In this case, the stress level reaches its steady state virtually instantly, which is: ˙η=αS(δ(S0)η)=0η=δ(S0).

    We are aware that having multiple environmental factors included into the model would make the advantages of using the concept of the environmental stress level more evident. However, due to the lack of corresponding data, we focus on the special case of having only one environmental variable (nutrient saturation S). We consider the investigations in this manuscript as a proof of principle that the environmental stress level is a feasible alternative way to model the effect of the environment on the cells.

    Overview over all models and their mathematical properties. The last model Mη has been constructed from model MS, which itself uses model Mopt as a basis. Therefore, the number of variables and parameters increases with adding more complex dynamics. An overview over the variables and strictly positive model parameters of each system as well as abbreviating notations for important parameter terms/functions can be found in Tables 1, 2 and 3 below. The units of the variables and parameters are motivated by the biological setting and the measurement methods, which are explained in more detail in the following Section 2.2.

    Table 1.  Overview over the model variables and their occurrence in the models.
    Mopt MS Mη Variable Meaning Unit
    V=V(t) Density of viable tumor cells 105 cells/ml
    S=S(t) Nutrient saturation 10% FBS
    η=η(t) Environmental stress level

     | Show Table
    DownLoad: CSV
    Table 2.  Model parameters, their occurrence in the models and their meaning.
    Mopt MS Mη Parameter Meaning Unit
    β Maximal possible proliferation rate 1/day
    K Carrying capacity of the biological system 105 cells/ml
    1/m Strength of proliferation contact inhibition
    λ Natural death rate 1/day
    λst Maximal possible starvation rate 1/day
    Sthr Nutrient sensitivity threshold 10% FBS
    αS Sensitivity rate of nutrient changes on stress level 1/day

     | Show Table
    DownLoad: CSV
    Table 3.  Abbreviating notations for important/useful parameter terms and functions as well as their occurrence in the model equations (none of them occurs in model Mopt).
    MS Mη Notation Meaning
    δ+(S0)= S02Sthr2+S02 Influence function for nutrient-promoted dynamics
    δ(S0)= 1δ+(S0) Influence function for nutrient-inhibited dynamics
    βS= δ+(S0)β Nutrient-dependent net proliferation rate (constant)
    λS= λ+δ(S0)λst Nutrient-dependent net death rate (constant)
    βη(t)= (1η(t))β Stress-dependent net proliferation rate (varying in time)
    λη(t)= λ+η(t)λst Stress-dependent net death rate (varying in time)

     | Show Table
    DownLoad: CSV

    A mathematical analysis of the models yields positivity and boundedness of the solutions, which are important features for biological reasonableness. Table 4 summarizes the computed bounds as well as the steady states ˉV reps. (ˉV,ˉη)T and their stability. The corresponding calculations providing these results can be found in the supplementary Section 4.2.

    Table 4.  Bounds of the solutions V(t), η(t) and steady states ˉV, (ˉV,ˉη)T of each ODE model.
    Positivity & boundedness Steady states
    Model of the solutions stable unstable
    Mopt 0V(t)max{V0,Km1λSβS} Km1λSβS 0
    MS for λS<βS: 0V(t)max{V0,Km1λSβS} Km1λSβS 0
    for λSβS: 0V(t)V0 0
    Mη for all λS,βS: 0η(t)max{η0,δ(S0)}1
    for λS<βS: 0V(t)max{V0,Km1λSβS} (Km1λSβS,δ(S0))T (0,δ(S0))T
    for λSβS: 0V(t)V0 (0,δ(S0))T

     | Show Table
    DownLoad: CSV

    We use data from the experiments described in [7] to calibrate the unknown parameters from Table 2. A CellTiter-Blue® assay was used to monitor the viability of tumor cells. In particular, viable cells metabolize a provided chemical and they emit measurable light as a result of this process. Hence, the data points are fluorescence intensity measurements. This way, viability was measured once every day and there are four biological replicates of each measurement to estimate statistical significance and repeatability.

    Let Itotal be an intensity measurement of a specific cell line at time point ˜t. Excluding the corresponding background intensity IBG of the cell-free medium, the fluorescence intensity produced by viable cells IV is assumed to be directly proportional to the density of viable tumor cells V(˜t). The experiments were performed with different initial cell densities between 104 and 105 cells per milliliter, which motivates the unit of V and eventually leads to the relation

    IV=ItotalIBGV(˜t)105 cells mln=IVV(˜t)(0,1),

    where n denotes the proportionality constant translating fluorescence intensity to cell density. Whenever we refer to "intensity measurements" in the following, we mean the fluorescence produced by the cells IV and neglect the superscript "V" for better readability.

    For nutrition the cells are supplemented with a particular concentration between 0% to 10% of fetal bovine serum (FBS). A supplementation with 0% FBS does not provide the cells with any nutrients, whereas 0% FBS generates optimal growth conditions. The nutrient supply is kept constant throughout the whole duration of each experiment, i.e. the nutrient saturation is assumed to stay at its initial level S0[0,1] at any time. Overall, this motivates:

    S(t)=S(0)=S0t0with unit[S]=0%FBS.

    The remaining variable, the environmental stress level η=η(t), is an auxiliary variable and especially immeasurable. Hence, it has no experimental counterpart and is used as a dimensionless quantity.

    How are uncertainties considered? In reality, the equation n=I/V(˜t) is not rigorously fulfilled. This can be due to e.g. model inadequacy or biological fluctuations of the cells' metabolism, which affect measurement accuracy. To capture this uncertainty, we assume a multiplicative noise for each element of a set of MN measurements {Ii}Mi=1 and a set of model solutions {Vi}Mi=1 considering the corresponding values of t, V0, and S0. In the Bayesian framework that we adopt (see Section 2.3) we model this to be a random variable εi such that for i=1,,M:

    Ii=nViεi. (2.5)

    Let εi be i.i.d. and have the unimodal and continuous distribution of a random variable ε with probability density function (PDF) fε. Then, the following properties should hold:

    [P1]:supp(fε)R+,i.e. all measurements are positive;[P2]:E[ε]=1,i.e. measurements are accurate on average;[P3]:fε(x)x0 and fε(x)x00,i.e. outliers are possible but not likely.

    Different distributions are possible to accomplish these properties. A small number of shape parameters and an easy calculation to ensure property [P2] are desirable. Therefore, we choose a Gamma distribution εΓ(a,b), a,b>0 with a few restrictions. This is a plausible choice for multiplicative noise and often used in imaging theory (see e.g. [21,22,23]). Property [P1] and the desired behavior near infinity of [P3] are satisfied by definition. The corresponding PDF is given by fa,bε(x)=baΓ(a)xa1ebx, where Γ() denotes the Gamma function. To fulfill the remaining properties, we observe

    E[ε]=ab [P2]=1, if b=alimx0fa,aε(x)=limx0(aaΓ(a)xa1eax) [P3]=0, if a>1.

    Note that by constraining b=a>1 the shape of the distribution depends only on the parameter a. In fact, a is directly related to the standard deviation σ and hence the variance of the distribution: σ2=Var(ε)=aa2=1aa=1σ2. Therefore, for i=1,,M the uncertainty factor εi for a particular measurement Ii can be modeled by

    εi=IinViΓ(1σ2,1σ2) with σ2(0,1). (2.6)

    We use the percentiles P5% and P95% of the Gamma distributed uncertainty factors εi to define the

    "90% (uncertainty) range" around the solution V:[VP5%,VP95%]. (2.7)

    In particular, given a specific noise variance σ2, the model expects 90% of the measurements within this interval, whereas respectively 5% are expected below and above it. The left side of Figure 2 depicts exemplary plots of the PDF fε for different values of σ2. It also shows the positive skewness of the distribution. Under consideration of the measurement method, this is a reasonable feature for the uncertainty factors, assuming the cells might not metabolize the assay to their full potential. On the right side of Figure 2 we see an example of a 90% uncertainty range around a solution.

    Figure 2.  Left: Probability density function fε of εΓ(1/σ2,1/σ2) for varying variances σ2 and the percentiles P5% and P95% for σ2=0.05. Right: Resulting 90% uncertainty range (shaded area) around an exemplary model solution V(t).

    The reason for considering a multiplicative (and, for instance, not additive) noise term is twofold. From a mathematical perspective, it allows to preserve positivity of the data. A more practical motivation is the reasonable assumption that the fluorescence noise of the intensity measurements is proportional to the density of viable cells. This is also supported by the observation of a larger variance for experiments with larger cell numbers in our data. Furthermore, multiplicative noise has demonstrated to be better suited for fluorescence than additive noise in other experimental settings [24].

    How can the measurements be utilized? One of the experiments in [7] monitors the dependence between cell viability and nutrient supply by providing five different nutrient concentrations S0 over a period of 7 days: this results in five data sets "D1"–"D5". These measurements are employed to calibrate the unknown model parameters for models MS and Mη with Bayesian inversion methods (for details see following Section 2.3). Another experiment in [7] was designed to investigate cell behavior under optimal growth conditions, i.e. the cells were provided with 10% FBS over a period of 21 days. We denote the data set generated by this experiment with "D6". It is used additionally to D1–D5 to validate the calibration results with model Mopt. All experiments start with several populations of different initial size V0. Table 5 shows the experimental setup of all data sets and the resulting initial conditions in the models.

    Table 5.  Overview over the data sets and the corresponding initial values. D1–D5 are used to calibrate the models MS and Mη and D6 for validation of the results with model Mopt.
    Nutrition Duration Initial values in the models
    Data set (in % FBS) (in days) V0 S0 η0
    D1 10.0 7 1.00,0.50,0.25 1.00 0.00
    D2 4.5 7 1.00,0.50,0.25 0.75 0.00
    D3 5.0 7 1.00,0.50,0.25 0.50 0.00
    D4 2.5 7 1.00,0.50,0.25 0.25 0.00
    D5 0.0 7 1.00,0.50,0.25 0.00 0.00
    D6 10.0 21 1.00,0.50,0.25,0.10,0.05 (1.00) (0.00)

     | Show Table
    DownLoad: CSV

    The task to identify the unknown true parameters from given data is called the "inverse problem". We solve it using a Bayesian approach, which leads, under mild assumptions, to a naturally well-posed inverse problem [25]. Furthermore, this approach allows to quantify the uncertainty in the estimated parameters and hence it brings more information than a deterministic inversion method, at the price of a higher computational cost.

    We collect all parameters to be estimated in a vector θΘRd (dN), where Θ denotes the parameter space. Assuming the considered data consist of M intensity measurements (excluding background intensity), we collect them in a vector I=(I1,,IM)T=(Ii)Mi=1RM. Defining G:ΘRM with θ(Gi(θ))Mi=1 as the forward operator mapping parameter values to the corresponding intensities, such a measurement can be rewritten as

    Ii(2.5)=nViεi=Gi(θ)εi, (2.8)

    where Vi is the corresponding model solution to (Mopt), (MS) or (Mη) using θ as parameters and εi is the multiplicative noise. Note that we use parts of the data sets D1–D5 for parameter estimation, hence each measurement Ii can refer to a different nutrient condition, initial cell density, and time point. Therefore, Vi has to be calculated in consideration of the corresponding values for S0, V0 and t.

    Now, we want to consider all measurements in I=(Ii)Mi=1 collectively. The parameter vector θ and the noise (εi)Mi=1 are modeled as multi-dimensional random variables taking values in Θ and RM respectively. The Bayesian formulation of the problem is the following: Given a prior (measure) μ0 on Θ, compute the posterior (measure) μI given the data I. The prior in the Bayesian setting is the correspondent of a regularization in deterministic inverse problems [25] and it reflects the knowledge about the parameters before including any information given by the data, whereas the posterior describes the knowledge after seeing the data. Let π0 and πI denote the probability densities of μ0 and μI, respectively. By Bayes' formula, we have

    πI(θ)=L(I|θ)π0(θ)ΘL(I|θ)π0(θ)dθL(I|θ)π0(θ), (2.9)

    where L is the data likelihood [25]. The proportionality constant of relation (2.9) depends only on I. It is called "model evidence" and it can be used to quantitatively compare two models (see Section 2.4). We remind from the previous Section 2.2 that (2.6) states i.i.d. εiΓ(1σ2,1σ2) with σ2(0,1) for every measurement Ii, i=1,,M. Using this together with (2.8), the data likelihood of I=(Ii)Mi=1 is:

    L(I|θ)Mi=1(IiGi(θ))1/σ21exp(1σ2IiGi(θ)).

    How to sample from the posterior? In order to make predictions, we want to sample from the given posterior distribution (2.9). However, this has a complicated, concentrated density, so we cannot sample from it exactly with a random number generator. To approximate the posterior measure, we use therefore the Sequential Monte Carlo (SMC) method, which we explain now based on [26]. In SMC one considers a sequence of intermediate distributions (μk)Nk=0, such that μ0 is the prior and μN=μI coincides with the posterior distribution. The probability density πk of the intermediate measure μk can be defined by

    πk(θ)=1Zkki=1Li(Ii|θ)π0(θ)or, equivalently, πk(θ)=1˜ZkLk(Ik|θ)πk1(θ), (2.10)

    where Zk and ˜Zk are normalizing constants and

    Li(Ii|θ)=1Zi(IiGi(θ))1/σ21exp(1σ2IiGi(θ))

    is the likelihood associated to observation Ii with normalization constant Zi, i=1,,M. The intermediate densities could also be constructed with an adaptive approach using tempering [9]. However, since the considered data measures the quantity of interest in a time series, the presented filtering method is computationally more efficient. The SMC algorithm samples sequentially from the intermediate measures μk using a weighted swarm of samples, called particles. Let PN be the sample size, i.e. the number of particles. At the k-th iteration (k=1,,N) the algorithm leads to a collection of particles {θp}Pp=1 with associated weights {Wkp}Pp=1, which gives the approximation

    πk(θ)Pp=1Wkpδθp(θ).

    The SMC algorithm is summarized in Algorithm 1: we achieve appropriately weighted particles to approximate μN=μI by starting with uniformly weighted particles distributed according to the prior μ0 (line 1) and iteratively move the samples from the 3-7) and a mutation step (line 8) [27,Ch. 5], which are explained in detail in the following paragraphs.

    Algorithm 1: Sequential Monte Carlo
    1: k=0: sample θpμ0 and set W0p=1P,p=1,,P
    2: for k=1,,N do
    3:   wkp=Lk(Ik|θ)Wk1p,Wkp=wkp/(Pp=1wkp),p=1,,P    ▷selection step
    4: if Peff<Pthr then resample:
    5: (i): sample indices acc. to distribution R of particle indices: (π1,,πP)R(Wk1,,WkP)
    6: (ii): set θp=θπp and Wkp=1P,p=1,,P
    7: end if
    8: move θpκk(θp,),p=1,,P   ▷μk-invariant mutation step
    9: end for

    Selection step. We start with a collection of particles {θp}Pp=1, distributed according to μk1. Their weights {Wk1p}Pp=1 are updated to {Wkp}Pp=1 by importance sampling: for p=1,,P we have

    Wkp=wkpPp=1wkpwithwkp(2.10)=Lk(Ik|θ)Wk1p.

    We see that Wkp is normalized to ensure that Pp=1Wkp=1, i.e. having a probability distribution. Note that the importance sampling only changes the weights and not the particles. However, if there are many particles with low weights, the estimation is only as accurate as a Monte Carlo approximation with a very small number of particles [28]. In this case, the reweighing step is followed by a resampling step, where the particles are replaced according to their updated weights. Resampling is needed, if the effective sample size

    Peff(Pp=1(Wkp)2)1

    is small, which we check by comparison with a threshold Pthr=τP,τ(0,1):

    Peff{<Pthr resample {θp}Pp=1 according to {Wkp}Pp=1 and uniformly weigh new particles,Pthr do not resample.

    This discards particles with low weight and improves the representation of the distribution μk.

    Mutation step. Performing only selection steps will eventually lead to degeneracy in the diversity of the particle population. In particular, after some resampling steps, few particles will survive and be replicated. Therefore, we introduce diversity in the particles by moving them according to a Markov Chain Monte Carlo (MCMC) kernel κk(,). This kernel is μk-invariant, i.e. it does not modify the particle distribution. We adopt the adaptive strategy developed in [9] to construct such a MCMC kernel. A random walk Metropolis-Hastings (MH) proposal is used on each univariate component, conditionally independently. More precisely, remembering that each particle is a vector θpRd, MH proposes {qp}Pp=1 with qpRd by computing

    (qp)j=(θp)j+ϵkjξjwithξjN(0,1),

    where ()j denotes the j-th component (j=1,,d) and the scale ϵkj is tuned to the acceptance rate of the previous SMC iteration. This means, we choose ϵkj adaptively as

    ϵkj=ρk^Var((θ)j)with recursively defined scaling parameterρk={ρk12if ak1>0.30,ρk1/2if ak1<0.15,ρk1otherwise,

    where ak1 is the average acceptance rate over the particles at the previous iteration and ^Var((θ)j) denotes the empirical marginal variance from the j-th components of all particles. It is also possible to adapt the proposal using empirical covariances instead [29], which would be computationally more expensive in view of possible extensions to parameters, which are random fields and therefore very high dimensional. Eventually, the particles are moved by randomly accepting the proposed ones:

    set θp=qp with probability min{πk(qp)πk(θp),1}.

    To improve the mixing, it is possible to repeat this process more than once by applying κk again on the moved particles. Afterwards, the total ratio of accepted particles determines the ak, which is used to scale ϵk+1j in the MCMC update of the next iteration.

    How is the theory applied to the models? Parameter estimation is performed in Python adapting the code provided in [28]. The set of parameters, which need to be calibrated, can be distinguished between model parameters (see Table 2) and hyper parameters. Latter includes the unknown uncertainty variance σ2 and the proportionality constant n from relation (2.6). All calibrations are performed using the SMC method with sample size P=50000 using the resampling threshold Pthr=75%P. In the mutation step, five MCMC updates are performed, i.e. line 8 in Algorithm 1 is performed five times.

    The parameters are calibrated over the course of N=83=24 SMC steps: As depicted on the left side of Figure 3, starting with the data corresponding to V0=1.00 and t=0, the SMC steps iterate over the eight points in time (inner loop) and over the three separate seeding densities (outer loop). With each step k, another set of 54=20 data points, containing all measurements of D1–D5 at a specific time point t regarding a particular initial cell density V0 (see right side of Figure 3), is included into the considered data collection Ik for the selection step (line 3 in Algorithm 1). This maximizes the information about the effect of varying nutrients on the cells at each calibration step. In particular, in the first step a set of measurements I1={I1,i}M1i=1 with M1=54=20 is considered and with each SMC step 20 more data points are included incrementally: I1I2IN with |Ik|=Mk=k20, k=1,,N. In total, the parameters are calibrated using MN=2420=480 measurements.

    Figure 3.  Schematic description of how the data sets D1–D5 are used to calibrate the parameters step by step with SMC.

    We want to construct informative prior distributions for the parameters. Information about the parameters' magnitude can be described by using an appropriate uniform distribution U(a,b) on an interval [a,b]. If we can additionally assume that the neighborhood of a parameter value h has a high probability, a triangular distribution Triang(a,h,b) on [a,b] is used, where h[a,b] is the mode of the distribution, i.e. the value that is most likely to be sampled. Note that the mode can also lie on the interval bounds, if low/large values within the interval are assumed to have high probability. We set h=(a+b)/2 to the center of the interval, if we expect the borders a,b to have a low probability but there is no particular tendency to a value within the interval.

    The doubling time T of the used cell type is larger than one day while in exponential growth [30,31], i.e. T>1. We can estimate an upper bound for the growth rate β:

    2V0=V(T)=V0eβT2=eβTβ=ln(2)TT>1<ln(2)<1.

    Therefore, it is reasonable to expect β(0,1). Under optimal nutrient conditions, the cell population size increases. This translates to the parameter relation β>λ, which can also be written as λ=c1β with c1(0,1). Since the parameter c1 is bounded, we calibrate this one instead of λ. A similar reparametrization can be done for the starvation rate λst\, : it is reasonable to assume that the cells die faster from starvation than from natural causes in case of a nutrient-free environment, which leads to λst>λλst=λc2=c1c2β with c2(0,1). Furthermore, the underlying experimental setting motivates the assumption that the population size does not exceed 3×105 cells per milliliter. All cell lines are seeded in initial densities below the carrying capacity and V0=1 is the largest seeding density, i.e. K(1,3) is plausible. The restrictions m>1 and Sthr(0,1) with Sthr1 are motivated by the modeling framework. Utilizing all this information, we adopt the following prior distributions for the model parameters of (MS):

    βU(0,1),c1Triang(0,1/2,1),c2Triang(0,1/2,1),KU(1,3),mU(1,12),  and  SthrTriang(0,0,1).

    The same prior distributions are used for model 2.4. Its equations only have one additional model parameter αS, for which we do not have any particular information – we set its prior distribution to

    αSU(0,12).

    To consider a certain degree of confidence in the measurements, we assume a small uncertainty variance σ2(0,1) is more likely than a large one. Therefore, we use Triang(0,0,1/2) as its prior. All experiments are started with cells from a batch with optimal nutrient conditions. If they are put into a nutrient-free environment without going through a weaning process beforehand, they can undergo a starvation shock. This might disturb or decrease the cells' ability to metabolize the chemical for the fluorescence measurements. To consider this in the hyper parameters, we allow the data sets D1–D4 (S0>0) and D5 (S0=0) to have different uncertainty variances σ2D1:4 resp. σ2D5 and proportionality constants nD1:4 resp. nD5, where potentially nD1:4nD5. Using the reparametrization nD5=cnnD1:4 with cn(0,1), leads to the prior distributions

    σ2D1:4,σ2D5Triang(0,0,1/2),nD1:4U(0,1/2),andcnTriang(0,1,1).

    A large value for the uncertainty variance σ2 allows a larger deviation of the model solution from the data. For the purpose of model comparison, using the same uncertainty variance for both models is desirable to increase comparability. Hence, we first calibrate each model separately to get an estimate for their variances σ2D1:4 and σ2D5. Then, we take the average ˉσ2 of the means respectively over both models, i.e.

    ˉσ2D1:4=12(ES[σ2D1:4]+Eη[σ2D1:4])andˉσ2D5=12(ES[σ2D5]+Eη[σ2D5]), (2.11)

    where "S" resp. "η" in the subscripts of the expected value E indicate the underlying model MS resp. Mη, which is used for the calibration. These average values ˉσ2D1:4 and ˉσ2D5 are then used deterministically and are especially not estimated anymore in further calibrations. For better comparability, we start each SMC algorithm from the same prior particle sample. In particular, the algorithm is performed for model Mη and the generated initial sample but without the component regarding to parameter αS, is used to start the calibration of model MS.

    The calibration results of each model are compared in a quantitative manner. For this, we calculate different comparison measures and validate the model solutions with the data sets D1–D6. An overview over the applied methods is given in the following paragraphs of this section.

    How can the quality of fit be quantified? We use the validation metric proposed in [32] to compare the model prediction at a given point in time with the corresponding set of measurements. Their mismatch is measured as the area between the data distribution Fdata and the prediction distribution FsolM using the calibration results of model M, mathematically defined by the metric

    d(Fdata,FsolM)=0|Fdata(I)FsolM(I)|dI. (2.12)

    For a set {Ii}i=1,,M of M intensity measurements, the data distribution function is given by

    Fdata(I)=Mi=1I(Ii,I)withI(Ii,I)={1/Mfor IiI,0for Ii>I.

    The prediction distribution function FsolM is determined approximately: for each weighted particle θp of the posterior, we calculate the empirical cumulative distribution function from the solution of model M. We get a set {˜Ip}p=1,,P with ˜Ip=npVp, where Vp is the model solution using the parameter sample θp scaled with the corresponding sample np of the proportionality constant. The prediction distribution is then

    FsolM(I)Pp=1IW(˜Ip,I)withIW(˜Ip,I)={WNpfor ˜IpI,0for ˜Ip>I,

    where WNp is the final weight of the p-th particle after the SMC algorithm is finished. Figure 4 depicts an exemplary plot of the above distribution functions and the resulting validation metric for a set of four measurements.

    Figure 4.  Plots of the data distribution function Fdata given by four exemplary intensity measurements I1,,I4, an approximate prediction distribution function FsolM and the corresponding validation metric d(Fdata,FsolM) given by the enclosed area between the two functions.

    How can the different model approaches be compared? We use the Bayes factor [33] for a quantitative model comparison. It gives the ratio between the respective model evidences. Latter describe the posterior probability of the data given the model type in consideration of the parameter distribution. For a model M, its evidence is defined by the marginal likelihood of a set of measurements I:

    p(I|M)=Θπ0(θ|M)L(I|θ,M)dθ,

    where π0(θ|M) denotes the prior on θ using model M and L(I|M,θ) the likelihood of I given M and its parameters θ. Following [10], the evidence Zk at the k-th SMC step can be approximated by

    ZkZ0kj=1Pp=1Lj(Ij|θp,M)Lj1(Ij1|θp,M)Wj1p,

    where P is the number of particles, Lj(Ij|θp,M) is the likelihood of observing all data sets included until calibration step j given the parameter sample θp with model M, and Wj1p denotes the normalized weight of θp at step j1. Since we start the calibrations with priors, from which we can sample directly, the initial evidence is Z0=1. The Bayes factor of two models M1 and M2 regarding the same set of measurements I is then given as

    Z1:2=p(I|M1)p(I|M2).

    For p(I|M1)>p(I|M2), i.e. Z1:2>1=100, the strength of evidence can be described by the following scale [10]:

    log10(Z1:2){(0,12]barely worth mentioning,(12,1]substantial support for M1,(1,2]strong support for M1,(2,)decisive support for M1. (2.13)

    This chapter investigates the results of the model calibrations. We compare the corresponding estimated parameters of both models MS and Mη and quantify the quality of fit to the data (Section 3.1). The calibration results can be used to identify parameter correlations (Section 3.2) and to study the reaction of the cells to different nutrient concentrations (Section 3.3). Additionally to the calibration data sets, the models are validated with further data (Section 3.4).

    To investigate dispersion of the results, we perform 12 runs of the SMC algorithm. In the following, the numerical deviations are given in terms of the mean and the 95% prediction/confidence interval of a normal distribution N(μ,σ2), i.e. denoted by μ±1.96σ. If not indicated otherwise, this formula is also used for error bars in visual representations of the deviations.

    According to the equations in (2.11), a precalibration of the models yields the uncertainty variances

    ˉσ2D1:4=0.0355andˉσ2D5=0.2410,

    which are used deterministically for further calibrations. We see that for the data sets D1–D4 (S0>0) the variance is considerably smaller than for data set D5 (S0=0). On the one hand, this could be a consequence of the previously described starvation shock, which might disturb the cells' metabolism. On the other hand, data set D5 contains mainly measurements of low cell density, for which the measurement accuracy might be decreased because of weak fluorescence.

    Regarding the estimated posteriors, the ESS is larger than Pthr after performing the last SMC step for all model calibrations. This implies that the final approximated posterior distribution is nearly as accurate as sampling directly from the correct probability measure. We calculate the expected values and variances (see Tables 6 and 7) of the model parameters' marginal posteriors.

    Table 6.  Expected values E() of the posterior distributions for each model parameter.
    Calibrated model parameters
    Model β λ λst αS
    MS 0.435±0.043 0.103±0.042 0.186±0.031
    Mη 0.437±0.023 0.106±0.025 0.196±0.018 6.930±2.561
    K m Sthr
    MS 1.740±0.133 4.731±2.371 0.104±0.011
    Mη 1.731±0.098 5.315±2.964 0.106±0.007

     | Show Table
    DownLoad: CSV
    Table 7.  Variances Var() of the posterior distributions for each model parameter.
    Calibrated model parameters
    Model β λ λst αS
    MS 0.007±0.017 0.006±0.014 0.010±0.020
    Mη 0.010±0.022 0.009±0.020 0.012±0.024 3.211±3.014
    K m Sthr
    MS 0.028±0.061 0.751±0.777 0.004±0.009
    Mη 0.033±0.067 1.071±1.782 0.004±0.008

     | Show Table
    DownLoad: CSV

    The tables show very similar values for both models and the differences between them lie within the range of the numerical variations of the SMC algorithm. The resulting average parameter values can be concluded as biologically reasonable, since we chose a prior distribution ensuring this. Up to some adaptions regarding units and reparametrization, the reasonableness of the parameters is also supported by comparing their values to the calibration results of [7], where a set of similar models is used together with parts of the same data.

    Taking a closer look at the nutrient sensitivity rate αS, we observe a large average variance of its marginal posterior (Var(αS)3.211) as well as large numerical deviations of its expected value (±2.561) and variance (±3.014). A possible reason for these large values is that only a minority of the measurements contain meaningful information to estimate αS. On the one hand, this parameter influences the cell growth mainly for early measurements, when the cells react to the nutrient change from the batch colony (S=1) to the provided nutrient supply of the respective experiment (S=S0[0,1]). On the other hand, the experiments with S0Sthr show little measurable reaction of the cells to nutrient changes due to the small value of the nutrient threshold Sthr, and therefore lack information to estimate αS. Overall, approximately less than 10% of the measurements contain meaningful information to estimate αS, which leads to a higher uncertainty in its estimation. Nevertheless, we can argue in Section 3.3 that this does not have a big influence on the solution of model 2.4.

    We investigate the ratio of the resulting values of the validation metric (2.12) of each model. This does not show any preference of a particular model (see Table 10 in supplementary Section 4.3). To get a quantitative statement for the models' quality of fit to the data, we calculate the logarithm of the Bayes factor: log10(Zη:S)=log10(p(I|Mη)p(I|MS)). In Figure 5 the average trend of this value over the course of model calibration steps is shown and interpreted according to the scale given in (2.13). Additionally to the averaged trend, the plot shows the evolution of the Bayes factor for one particular run of the algorithm (dashed line). For this run, the resulting expected values of the parameters' posteriors match approximately with the ones in Table 6. We will refer to the corresponding posteriors later in Section 3.2, when we investigate the correlation between the model parameters.

    Table 8.  Percentage of data points in D1–D5, which are below the 90% uncertainty range for different initial cell densities V0 (rows) and days t (columns). If there are numerical deviations between the runs, they are given in terms of the 95% confidence interval of a normal distribution.
    Day of measurement (i.e. t)
    Model V0 0 1 2 3 4 5 6 7
    MS 1.00 0 0.8±3.8 0 0 0 0 0 0
    0.50 5.8±3.8 0 5.8±3.8 35.0 0 0 0 0
    0.25 13.8±6.1 5.0 30.8±3.8 73.8±4.4 5.4±2.8 0 1.3±4.4 4.2±3.8
    Mη 1.00 0 0 0.8±3.8 5.0 3.3±4.8 0 5.0 5.0
    0.50 5.0 0 4.9±5.0 35.0 0 0 0 0
    0.25 14.2±3.8 5.0 32.1±5.0 72.9±5.0 5.4±2.8 0 1.3±4.4 5.0

     | Show Table
    DownLoad: CSV
    Table 9.  Mean values and variances of the posterior distributions for some calibrated parameters. The deviations between the runs of the SMC algorithm are given in terms of the 95% prediction interval of a normal distribution N(μ,σ2): μ±1.96σ.
    Calibrated parameters
    Model nD1:4 nD5 cn c1 c2
    MS E() 0.244±0.006 0.190±0.020 0.780±0.084 0.236±0.076 0.576±0.293
    Mη 0.243±0.004 0.182±0.022 0.752±0.094 0.240±0.046 0.554±0.159
    MS Var() 0.002±0.004 0.008±0.017 0.034±0.069 0.011±0.024 0.053±0.100
    Mη 0.002±0.004 0.007±0.014 0.032±0.055 0.016±0.036 0.069±0.122

     | Show Table
    DownLoad: CSV
    Table 10.  Ratio dη/dS of the validation metrics using model Mη resp. MS averaged over all time points considering measurements regarding various V0 (columns) and data sets, i.e. different S0 (rows). The values in the last row/column are averaged over all data sets/initial cell densities before taking the mean over the SMC runs. The numerical deviations between the runs are given in terms of the 95% confidence interval of a normal distribution.
    Initial cell densityV0
    Data set 1.00 0.50 0.25 0.10 0.05 all (avg.)
    D6 1.05±0.54 1.04±0.33 1.07±0.17 1.01±0.24 1.03±0.13 1.04±0.26
    D1 1.04±0.14 1.01±0.12 1.00±0.09 - - 1.02±0.04
    D2 1.01±0.07 0.99±0.17 0.99±0.15 - - 1.00±0.12
    D3 1.02±0.04 1.02±0.21 1.03±0.13 - - 1.02±0.07
    D4 1.00±0.06 1.01±0.13 1.01±0.04 - - 1.01±0.07
    D5 0.94±0.15 1.06±0.18 1.03±0.08 - - 1.01±0.11
    all (avg.) 1.01±0.09 1.02±0.12 1.02±0.04 1.01±0.24 1.03±0.13 1.02±0.09

     | Show Table
    DownLoad: CSV
    Figure 5.  Logarithm of the Bayes factor Zη:S over the course of the SMC steps, where the error bars indicate the 95% confidence interval resulting from the various runs of the SMC algorithm. The dashed trendline shows the evolution for a single run, which estimated the model parameters close to the average values of all runs. The gray labels on top of the plot give the included data points at each SMC step.

    For the first eight calibration steps, the Bayes factor indicates strong evidence to prefer model Mη over MS. Until this point only data regarding V0=1.00 are considered. Incremental inclusion of the next data set (V0=0.50) weakens the weight of the support for model Mη, but still allows to conclude a tendency of the evidence towards this model. This can be observed further until SMC step 19. At step 20 (i.e. with inclusion of data with t3 and V0=0.25), the uncertainty suddenly increases drastically, which does not allow a clear interpretation of the Bayes factor anymore. The 95% confidence interval reaches from areas of decisive support for model Mη to decisive support for model MS. After this step the uncertainty decreases again but stays on a high level, still not allowing a clear interpretation of the Bayes factor. Overall, the average trend of the Bayes factor indicates support for model Mη, despite needing an additional variable and parameter compared to model MS, which is already implicitly penalized by the Bayes factor [33]. The modeling advantages of using the environmental stress level presented in Section 2.1 still hold true, even though the present experimental data are not best-suited for showing quantitatively in a prominent way the benefits of this approach. Further investigations with better-suited data are needed to demonstrate the full potential of the environmental stress level.

    To understand the large deviations in the Bayes factor, especially the one at SMC step 20, we take a closer look at the model solution and the data. We use the average expected values from Table 6 (model parameters) and supplementary Table 9 (scaling factors nD1:4 and nD5) to calculate the corresponding average model solution V(t) and scale the measurements IV. The resulting time evolution of V(t) in comparison with the calibration data sets D1–D5 can be seen in Figure 6.

    Figure 6.  Time evolution of the average model solution V of model MS resp. Mη for different V0 compared to the corresponding median of the scaled measurements IV/n. The error bars show the median absolute deviation considering the four biological replicates.

    In general, we observe a good fit for both models. As expected, the similar model parameter values in Table 6 result in nearly the same solution for each model – at least for all cases of S0>0 (first two rows of plots). In these cases the estimated proportionality constant is nD1:40.24 for both models, whereas for S0=0 they differ: nD50.19 resp. nD50.18 (MS resp. Mη). This results in different scaling of data set D5 (bottom row of Figure 6). However, the difference of the value of nD5 between the models lies in the scale of numerical deviations of the SMC algorithm (see supplemented Table 9).

    Taking a closer look into the measurements in Figure 6, the measured population size drops for all S0 at the third day of the experiment (t3). This measurement might be an outlier and it corresponds to the data included in SMC step 20, where we observe a large deviation in the Bayes factor. To investigate if this drop influences the quality of the fit, we consider the underlying measurement uncertainty. Dependent on the uncertainty variance σ2, we can calculate the percentiles P5% and P95% of the Gamma distributed uncertainty factor ε:

    ˉσ2D1:4=0.0355P5%=0.712,P95%=1.329,ˉσ2D5=0.2410P5%=0.350,P95%=1.920.

    These determine the 90% uncertainty range (2.7) for the calibration data D1–D4 resp. D5, i.e. give an interval around the solution V(t), where the model expects 90% of the measurements.

    We want to know how many measurements are actually situated where they are expected. For each data set, we count the scaled measurements IV/n, which are situated below/within/above the 90% uncertainty range of the corresponding average solution V. A graphic overview over the calculated percentages can be found in the supplementary Section 4.3 (Figure 10). We see that averaging over the data sets, for both models roughly 90% of the scaled data is actually situated within the 90% uncertainty range. Regarding the remaining data points, we observe that the model solutions tend to be larger than the measurements, since about 8.3–8.5% of the data points lie below the 90% range, whereas only about 2.1% are above. To investigate this observation, we take a closer look at the data below the 90% range by checking if the discrepancy focuses on specific measurements, see Table 8.

    Figure 10.  Percentage of data points which are situated below/within/above the 90% range [VP5%,VP95%] around the solution V. We remind that the error bars indicate the numerical deviation between the SMC runs. The dotted lines in the corresponding color give the average over all data sets D1–D5.

    On the third day, the measurements show an extraordinarily large amount of data below the 90% uncertainty range, where only 5% would be expected. This supports the hypothesis that the measurements on that day might be outliers. The percentages are higher the smaller V0 is: approx. 35% resp. 68–78% of the measurements are below the 90% range for V0=0.50 resp. V0=0.25. This negative correlation is expectable, since the width of the range [VP5%,VP95%] decreases with smaller V (see right side of Figure 2). This also explains another observation: within the measurements of a particular day, more data points tend to be below the 90% range for smaller V0. Overall, these discrepancies between measurements and expectation can be a reason for the high deviations of the Bayes factor when including data with V00.50 in the SMC calibration (see Figure 5). In particular, the sudden increase in deviation at SMC step 20 can be linked to the potential outliers for t3 and V0=0.25.

    The decrease of population size on the third day can be observed for almost all seeding densities V0 and nutrient saturations S0>0, although there is no biological reason for this as long as there are nutrients available. Hence, it is reasonable to assume that the discrepancies result from an unknown experimental bias rather than modeling inaccuracy or biological variation.

    To identify parameter correlations, we investigate the estimated posteriors from a single run of the SMC algorithm. We choose a run whose resulting expected values of the parameters' posteriors match approximately the ones in Table 6. Figure 7 shows the corresponding marginal posteriors of the model parameters for each model (first row/column) and samples drawn from the 2D posterior distributions of pairwise parameter combinations. The shape of the scatter plots can give an impression regarding the correlation of parameters.

    Figure 7.  The horizontal/vertical axis labels are given on the bottom/right side. Plots in first row/column: Marginal posterior distributions of the model parameters using model MS resp. Mη (note: magnitude of KDE axis is neglected and not the same for the different parameters). Remaining plots: Scatter plots for 5000 samples drawn from the 2D distributions of corresponding pairwise parameter combinations (red, below diagonal: MS; blue, above diagonal: Mη). A regression line and correlation coefficient r is depicted if a linear correlation is observable.

    It is observable that the SMC algorithm leads to unimodal distributions. The scatter plots indicate no correlation for most of the parameters, as the points are not concentrated along a curve. In particular, the stress-related parameter αS does not correlate with other parameters (last column of scatter plots). For both models, a linear correlation can be observed between β and λ (MS: r0.84; Mη: r0.89) as well as between K and m (MS: r0.68; Mη: r0.58). These correlations are statistically significant, since their p-values are sufficiently small (0.01).

    We have seen in Section 2.1 that the ratio between the growth rate β and the natural death rate λ is crucial for the behavior of the population: it determines the "net" growth rate ˜β=βλ and the "net" capacity ˜K=K(1λ/β)1/m. Since ˜β,˜K>0 should hold, those quantities require the relation β>λ. This motivates the observed strong positive correlation (r>0.7). For the carrying capacity K and the parameter m, we see a moderate negative correlation (0.7<r<0.5). These parameters regulate the proliferation contact inhibition in the ODEs of both models MS and Mη with the term (V/K)m. This term is smaller, the larger K or m are. Hence, these parameters need to be negatively correlated to mathematically describe a certain level of contact inhibition.

    To affect cell growth and death, both models use the same influence functions δ±(S). In model MS these functions scale the rates β and λst directly, whereas in model Mη they scale them indirectly via the stress level η(t). Therefore, the value of the nutrient sensitivity threshold Sthr should not depend on the choice of the model. This behavior is actually observable in the results: The expected value of the estimated nutrient sensitivity threshold Sthr appears to be nearly identical for both models with E[Sthr]0.1 (see Table 6), resulting in the same shape of the influence function δ(S).

    We use the corresponding expected values from Table 6 to calculate the stress level in model Mη. The resulting time evolution of the stress level η(t) is depicted in the first two plots of Figure 8. The left plot shows η(t) for fixed Sthr and varying αS, while the middle one also fixes αS to its calibrated mean. Fixing Sthr is feasible, as we concluded in previous Section 3.2 that it is uncorrelated to αS. The third plot shows the corresponding influence functions δ±(S).

    Figure 8.  Time evolution of the environmental stress level η(t) with Sthr=0.106 for different nutrient conditions S0 (left plot: αSN(6.930,2.567/1.96); middle plot: αS=6.930), where its steady states coincide with the corresponding values δ(S0) of the influence function (right plot).

    Since the estimated value of the sensitivity parameter αS is large, the steady state is already reached after approximately one day. This holds despite the relatively large deviations of the calibrated value for αS (see left plot). Therefore, variation of αS within the calibrated scale does not change the solution of model Mη significantly. As calculated in Section 2.1, the steady state of η(t) coincides with the function value δ(S0) of the influence function (see right plot). Hence, the high nutrient sensitivity explains why the calibrated model parameters and the resulting model solutions turn out to be very similar.

    For additional validation, we calculate the average solution of the "limit model" Mopt using the posteriors (determined with data sets D1–D5) to compare it with data set D6. In this experiment tumor cells were seeded in five different initial densities and supplied with 10% FBS for the duration of 21 days (see Table 5). Hence, it holds S0=1>0 (like in data set D1) and we use E[nD1:4]=0.24 to scale the measurements. The number of viable cells is calculated by inserting the estimated expected values of the model parameters β, K, m and λ from Table 6 into the analytical solution (2.1). Figure 9 compares the corresponding scaled data with the calculated time course of V(t).

    Figure 9.  Time evolution of the average model solution V of model Mopt for different V0 using the parameter estimates from the calibration of model MS resp. Mη compared to the corresponding median of the scaled measurements IV/0.24. The error bars show the median absolute deviation considering the four biological replicates.

    As expected, both solutions are very similar, as for model Mopt only the parameters β, K, m, and λ are relevant, which have similar estimates for both MS and Mη. Furthermore, we see a good fit to the data (see also ratios of validation metrics in Table 10 in the supplement). It is observable that the model solutions tend to overestimate the data after the steady state is reached. In reality, population growth often follows the concept of "overshoot": this describes the observation that a population might exceed its carrying capacity and drop back down afterwards (see e.g. [34]). Reasons can be a lag in sensing the available resources in the environment or the time interval between birth and death. The measurements starting with V0{1.00,0.50,0.25} (orange, purple, and green markers in Figure 9) suggest an observable overshoot effect, but the model solutions cannot reproduce this phenomenon due to the limited structure of model Mopt. This could explain the discrepancy between solution and data close to the carrying capacity.

    In summary, the presented results show that model Mη is an appropriate alternative way to describe the influences of nutrient changes on viable cells. The model preserves positivity and boundedness of the solutions and calibration yields biologically reasonable parameter values. Keeping in mind that large deviations from the data sets D1–D5 focus on a particular measurement (t3), we altogether see a good agreement to the data for different nutrient scenarios. For the simple setting with the nutrients as the only environmental variable, there is no strong need from a quantitative perspective to model its influence by using the additional variable η. But even in this special case, the Bayes factor indicates a preference for including the environmental stress level η into modeling. A mathematical model should be kept as simple as possible, but as additional influencing factors may become relevant, a suitable and flexible modeling interface is important. From a qualitative perspective, the general modeling approach (2.3) can provide such a setup.

    The introduction of an artificial stress level to describe different environmental factors in a collective manner has several advantages. Due to the modular structure of the general ODE for the stress level η (equation 2.3), this model is easily extendable to several environmental variables. This enables the consideration of different sensitivities of the cells to changes in specific environmental factors. If the values of the corresponding sensitivity rates α±j can be recovered from data, the magnitude of them can give insight into time scales of the cells' reaction to different environmental changes. This information can be used on the one hand to simplify the model by quasi-steady state assumptions for environmental variables regarding large sensitivity rates. On the other hand, if some sensitivity rates are estimated to be close to zero, the model can be reduced easily by setting them to zero and neglecting the ODEs of the corresponding environmental variables. The resulting quality of fit of the simplified model can then be investigated e.g. by the Bayes factor.

    Overall, the model allows to describe a survival strategy of the cells assuming they adjust their behavior to avoid a stressful environment by prioritizing according to their sensitivities. If spatial inhomogeneity is considered (i.e. if the model is extended to a system of partial differential equations), such an avoidance strategy can be modeled by influencing the cells' motility. This can be mathematically described similarly to chemotaxis: cells might try to move preferably to areas with a low stress level to improve their growth/survival conditions. Future work aims to investigate the application of the proposed general model in such a setting, especially considering mechanical environmental features like pressure or stiffness of the surrounding tissue, as considered experimentally e.g. in [35]. New experimental approaches like "organ-on-a-chip" will allow for even more realistic and nevertheless (as in vitro technique) well-controlled settings and corresponding parameter calibration [36,37].

    This paper was supported by Deutsche Forschungsgemeinschaft (DFG) through TUM International Graduate School of Science and Engineering (IGSSE), GSC 81.

    All authors declare no conflicts of interest in this paper.

    The following equations (4.1) and (4.3) are the case-dependent analytical solutions of system (MS) as given in Section 2.1. The corresponding calculations are done in the following paragraphs. It can be shown by differentiation that these derived terms for V(t) actually solve the given initial value problem.

    Case 1: βS>λS. Model MS has the same form as model Mopt just with nutrient-scaled growth rate βS=δ+(S0)β and death rate λS=λ+δ(S0)λst, each depending on the constant nutrient supply S0. Hence, for βS>λS the ODE of (MS) can again be written in the form of a logistic growth term and the analytical solution can be given analogously to equation (2.1) with scaled rates:

    V(t)=((V0K)m(βSλS)βSV0m+((βSλS)KmβSV0m)e(βSλS)mt)1/m. (4.1)

    Case 2: βS<λS. In case of βSλS the ODE cannot be rewritten in pure logistic form, since the root's argument of the resulting "net" carrying capacity K(1λS/βS)1/m might cause problems. However, it can be shown that for βS<λS the analytical solution above is still well-defined, since the root's argument is non-negative:

    V(t)(4.1)=((1)(1)0(V0K)m>0(λSβS)βSV0m0(1+e(λSβS)mt>1)+(λSβS)Km>0e(λSβS)mt>1)1/m. (4.2)

    Case 3: βS=λS. For λSβS the fraction inside the root of formula (4.2) degenerates, as numerator and denominator both tend to zero. Since V(t) is a differentiable function in the parameter space, we can derive the analytical solution in this case using l'Hospital's rule (l'H). It holds

    limλSβSV(t)ml'H=limλSβS(V0K)m0(λSβS)βSV0m(e(λSβS)mt1)0+(λSβS)0Kme(λSβS)mt1l'H=limλSβS(V0K)mβSV0mmte(λSβS)mt+Kme(λSβS)mt+(λSβS)Kmmte(λSβS)mtl'H=limλSβS(V0K)m(mt(βSV0m+(λSβS)0Km)+Km)e(λSβS)mt1=(V0K)mmtβSV0m+Km,yieldinglimλSβSV(t)l'H=V0K(mtβSV0m+Km)1/m. (4.3)

    In the following paragraphs, the solutions of all models are analyzed in terms of positivity and boundedness. Furthermore, we give the respective steady states and check for their stability.

    Positivity and boundedness of the solutions. In a reasonable biological context, the size of a cell population cannot grow arbitrarily because of environmental limitations. Hence, for each of the presented models to be feasible, the corresponding solution V(t) has to stay bounded and especially non-negative. The same properties are desired for the environmental stress level η.

    Environmental stress level. Before analyzing the variable V in all models, we start with the stress level η from system (2.7). Given η(0)=η0[0,1], its ODE yields

    ˙η=αS(δ(S0)η){>0for η<δ(S0)=0for η=δ(S0)<0for η>δ(S0), (4.4)

    leading to both positivity and boundedness of η: 0ηmax{η0,δ(S0)}1. Analogous bounds can be calculated for the general ODE (2.5) for the environmental stress level, since α±i,δ±j0.

    Density of viable cells. Regarding the remaining variable V, positivity and boundedness in model Mopt follow directly from the fact that its ODE can be rewritten as a logistic growth model with positive growth rate ˜β and capacity ˜K, preserving positivity and boundedness by definition. To conclude these properties also for models MS and Mη, we analyze a more general, non-autonomous model of the form

    {˙V=h(t)β=βh(t)V(1(VK)m)(λ+(1h(t))λst)=λh(t)V,V(0)=V0, (4.5)

    where h is a potentially time-dependent and bounded function: h(t)[0,1]t0. In fact, problem (4.5) is a generalization of both systems (MS) and (Mη), where respectively h(t)=δ+(S0) and h(t)=1η(t)(2.6)=1δ(S0)(1eαSt)η0eαSt. For h(t)=δ+(S0) the required boundedness of h is given by definition of δ+(S) and for h(t)=1η(t) it is given by the previously derived bounds: η[0,1]. Hence, we can analyze the general problem (4.5) for positivity and boundedness instead and transfer the results to models MS and Mη.

    With h(t)[0,1]t0 and V00, positivity follows directly by observing ˙V=00 for V=0. To analyze boundedness, we check the sign of the derivative of V. For λh(t)βh(t) we estimate:

    ˙V(4.5)=βh(t)V(1(VK)m)λh(t)V=βh(t)V(1λh(t)βh(t)0(VK)m)0.

    In case of λh(t)<βh(t), we can rewrite the latter term further by placing (1λh(t)βh(t)) outside the bracket:

    ˙V=(βh(t)λh(t))>0V(1(VK(1λh(t)/βh(t))1/m>0)m){>0for V<K(1λh(t)βh(t))1/m,=0for V=K(1λh(t)βh(t))1/m,<0for V>K(1λh(t)βh(t))1/m. (4.6)

    Together with βh(t)=h(t)ββ and λh(t)=λ+(1h(t))λstλ, this eventually leads to a time-independent upper bound for the population size:

    Vmax{V0,K(1λh(t)βh(t))1/m}max{V0,K(1λβ)1/m}.

    Note that relations (4.4) and (4.6) also fit to the non-trivial steady states of the corresponding models (see next paragraph).

    Steady states of the models. To calculate the steady states of each problem, we introduce the following notations for the right hand sides of the corresponding autonomous ODEs in each model:

    (Mopt):˙V=fopt(V),(MS):˙V=fS(V),(Mη):˙V=fη(V,η),˙η=gη(V,η). (4.7)

    The steady states are derived by setting the right hand sides to zero respectively. From a biological point of view, we would expect at least two steady states ˉV for the cell density. The cells can either go extinct or the population size tends to the "net" carrying capacity, which depends on the relation between growth and death rate. For a growing population, we use more general notations during the analysis: Let b be the total growth rate, which is larger than the total death rate a and define the function

    K:{(a,b)R2+|a<b}(0,K)with(a,b)K(1a/b)1/m

    to be the "net" carrying capacity depending on the rates. Then, the mentioned expected steady states translate to ˉV=0 or ˉV=K(a,b).

    System (Mopt). For model Mopt these are actually the only two steady states:

    ˙V|V=ˉV(Mopt)=βˉV(1(ˉVK)m)λˉV=ˉV(β(1(ˉVK)m)λ)=0ˉV=0orβ(1(ˉVK)m)λ=0β>λˉV=K(1λβ)1/m=K(λ,β).

    Note that the same non-trivial steady state can also be calculated by considering the limit t of the corresponding analytical solution (2.1).

    System (MS). Since model MS only differs from model Mopt by time-independent scaling of the growth and death rate, analogous calculations lead to a similar result:

    ˙V|V=ˉV(MS)=βSˉV(1(ˉVK)m)λSˉV=0ˉV={0for λSβS,0 or K(1λSβS)1/m=K(λS,βS)for λS<βS.

    Hence, we have the two expected steady states in case of λS<βS, but only the trivial one if λSβS.

    System (Mη). For the last model Mη, we have to consider a system of two ODEs:

    (Mη):{˙V|V=ˉV,η=ˉη=(1ˉη)βˉV(1(ˉVK)m)(λ+ˉηλst)ˉV=0,˙η|V=ˉV,η=ˉη=αSδ(S0)(1ˉη)αSˉηδ+(S0)=αS(δ(S0)ˉη)=0.

    Latter equation is independent of V, which directly results in the steady state for the stress level: ˉη=δ(S0). Inserting this into the first equation makes it identical to the equation from the previous paragraph for system (MS). Therefore, the steady states of problem (Mη) are

    (ˉVˉη)=(0δ(S0)) for all λS,βSand(ˉVˉη)=(K(1λSβS)1/mδ(S0))=(K(λS,βS)δ(S0)), if λS<βS.

    Stability of the steady states. The derived steady states can be analyzed for stability. We use the previously introduced notations (4.7) for the right hand sides of the ODEs. Following the theorem of Hartman and Grobman [38,39], local stability of a steady state can be evaluated by checking the sign of the derivatives of the right hand sides in these points.

    Systems (Mopt) and (MS). For the first two models we have to check the sign of fopt(ˉV) and fS(ˉV) respectively. Keeping the constraint β>λ in mind, inserting the steady states of model Mopt yields

    fopt(V)=βV(1(VK)m)λVfopt(V)=βλ(m+1)βKmVmresulting infopt(0)=βλ>0andfopt(K(1λβ)1/m)=βλ(m+1)(βλ)=m(βλ)<0.

    Hence, the non-trivial steady state is stable, whereas the trivial one is not. Since fS has the same form as fopt only with scaled growth and death rate, analogous calculations result in

    fS(0)=βSλS{>0for λS<βS,=0for λS=βS,<0for λS>βS.and, in case of λS<βS:fS(K(1λSβS)1/m)=m(βSλS)<0.

    The stability of the steady states now depends on the relation between the scaled growth rate βS and death rate λS. For the first case λS<βS, the trivial steady state is locally unstable, whereas the non-trivial one is stable. For λS>βS only the trivial steady state exists, which is stable.

    In case of equality λS=βS, the theorem of Hartman and Grobman cannot be used to determine stability for ˉV=0 and we take another approach by investigating the change of a time-dependent perturbation from the steady state ˉV=0. Observing that fS(V) is a polynomial in V of degree m+1 without an absolute term, all but one derivatives vanish in ˉV=0:

    f(j)S(0)={βSKm(m+1)!(m+1j)!0m+1j=0for j<m+1,βSKm(m+1)!=const.0for j=m+1,0for j>m+1. (4.8)

    Let ν=ν(t)>0 be the time-dependent perturbation from the steady state ˉV=0 with initial value ν(0)=ν0. The change of the perturbation in time can be derived by insertion into the ODE, i.e. fS(ˉV+ν(t))ODE=ddt(ˉV+ν(t))=ddtν, and Taylor series expansion of the right hand side:

    ddtν=fS(ˉV+ν(t))(4.8)=(mj=0f(j)S(0)=0(ν0)jj!)+(f(m+1)S(0)(m+1)!(ν0)m+1)+(j=m+2f(j)S(0)=0(ν0)jj!)=βSKmνm+1.

    The derived ODE in ν(t) can be analytically solved by separation of variables, resulting in

    ν(t)=ν0K(mtβSνm0+Km)1/m.

    Taking the limit t shows that the perturbation vanishes over time, i.e. local stability of the steady state ˉV=0 can be concluded.

    System (Mη). For the last model Mη we again use Hartman-Grobman theorem to investigate the corresponding locally linearized model. Reminding that the system is given by

    (Mη):{˙V=fη(V,η)=(1η)βV(1(V/K)m)(λ+ηλst)V,˙η=gη(V,η)=αSδ(S0)(1η)αSδ+(S0)η=αS(δ(S0)η),

    linearization in the steady state (ˉV,ˉη) with the Jacobian J yields

    (˙x˙y)=J(xy)=(VfηηfηVgηηgη)|(V,η)=(ˉV,ˉη)(xy)=((1ˉη)β(1(m+1)(ˉV/K)m)(λ+ˉηλst)βˉV(1(ˉV/K)m)λstˉV0αS)(xy).

    The eigenvalues χ1,2 of the triangular Jacobian J in dependence on the steady state (ˉV,ˉη) are given by

    χ1=αS<0andχ2=χ2(ˉV,ˉη)=(1ˉη)β(1(m+1)(ˉV/K)m)(λ+ˉηλst).

    The first eigenvalue χ1 is always negative, whereas the sign of χ2 depends on the relation between βS and λS. In particular, the non-trivial steady state only exists for λS<βS, leading to

    χ2(K(1λS/βS)1/m,δ(S0))=βS(1(m+1)(1λS/βS))λS=m(βSλS)<0,

    i.e. this steady state is a stable node. For the trivial steady state it holds

    χ2(0,δ(S0))=βSλS{>0for λS<βS,=0for λS=βS,<0for λS>βS,

    and, therefore, this steady state is a stable node for λS>βS and an unstable one for λS<βS. In case of equality λS=βS the second eigenvalue gets zero and the theorem is again not applicable anymore. To show stability in this case, we use the fact that the ODE of η is independent of V, i.e. gη(V,η)=gη(η), and thus the equations of model Mη can be analyzed consecutively. On the one hand, it holds

    gη(η)=αS(δ(S0)η)gη(η)=αS<0

    and on the other hand, inserting ˉη=δ(S0) into the first ODE of the system yields

    fη(V,ˉη)=βSV(1(V/K)m)λSV=fS(V).

    For this equation stability was already shown for model MS in case of λS=βS. Eventually, the trivial state of model Mη is stable for this case as well.

    Calibrated parameters. For the proportionality constants and reparametrization parameters, the expected values and variances of the posteriors are listed in Table 9.

    Uncertainties in the model. In Figure 10 we consider the 90% uncertainty range (2.7) with the percentiles P5% and P95% calculated in Section 3.1, using either the calibrated parameters of MS or Mη (left/right plot). There are three different markers per data set (vertical axis), showing how many data points (horizontal axis) are situated below/within/above the 90% range. For instance regarding D1 (topmost on vertical axis), in both plots the black bullet marker shows that approx. 87% of the measurements actually lie within the 90% range, whereas the dark/light blue triangle marker shows that about 4%/9% of them are above/below the 90% range.

    Quality of fit to the data. The validation metric (2.12) measures the mismatch between data and model prediction, therefore

    dηdS=d(Fdata,FsolMη)d(Fdata,FsolMS){1indicates: better fit for Mη,1indicates: MS andMηfit equally well,1indicates: better fit forMS.

    Calculations for each data set D1–D6 and initial cell number V0 result in the values given in Table 10. For data set D6 the solution of model Mopt was used by inserting the average estimated parameters resulting from the calibration of MS resp. Mη.

    All values are close to one and the differences between them are in the scale of numerical variations of the SMC algorithm. Hence, the investigation of the validation metric does not show any preference of a particular model.



    [1] L. Preziosi (ed.), Cancer Modelling and Simulation, 1st edition, CRC Press, 2003.
    [2] H. Byrne, T. Alarcon, M. Owen, S. Webb, P. Maini, Modelling aspects of cancer dynamics: A review, Philosoph. Transact. Ser. A Math. Phys. Eng. Sci., 364 (2006), 1563–1578. https://doi.org/10.1098/rsta.2006.1786 doi: 10.1098/rsta.2006.1786
    [3] K. Brindle, New approaches for imaging tumour responses to treatment, Nat. Rev. Cancer, 8 (2008), 94–104. https://doi.org/10.1038/nrc2289 doi: 10.1038/nrc2289
    [4] D. A. Hormuth, A. M. Jarrett, E. A. B. F. Lima, M. T. McKenna, D. T. Fuentes, T. E. Yankeelov Mechanism-based modeling of tumor growth and treatment response constrained by multiparametric imaging data, JCO Clin. Cancer Inform., 3 (2019), 1–10. https://doi.org/10.1200/CCI.18.00055 doi: 10.1200/CCI.18.00055
    [5] T. E. Yankeelov, R. G. Abramson, C. C. Quarles, Quantitative multimodality imaging in cancer research and therapy, Nat. Rev. Clin. Oncol., 11 (2014), 670–680. https://doi.org/10.1038/nrclinonc.2014.134 doi: 10.1038/nrclinonc.2014.134
    [6] C. S. Szot, C. F. Buchanan, J. W. Freeman, M. N. Rylander, 3D in vitro bioengineered tumors based on collagen I hydrogels, Biomaterials, 32 (2011), 7905–7912. https://doi.org/10.1016/j.biomaterials.2011.04.001 doi: 10.1016/j.biomaterials.2011.04.001
    [7] E. A. B. F. Lima, N. Ghousifam, A. Ozkan, J. T. Oden, A. Shahmoradi, M. N. Rylander, et al., Calibration of multi-parameter models of avascular tumor growth using time resolved microscopy data, Sci. Rep., 8 (2018), 1–14. https://doi.org/10.1038/s41598-018-32347-9 doi: 10.1038/s41598-018-32347-9
    [8] C. P. Robert, G. Casella, G. Casella, Monte Carlo statistical methods, 2nd edition, Springer-Verlag, New York, 2004. https://doi.org/10.1007/978-1-4757-4145-2
    [9] A. Beskos, A. Jasra, E. A. Muzaffer, A. M. Stuart, Sequential Monte Carlo methods for Bayesian elliptic inverse problems, Stat. Comput., 25 (2015), 727–734. https://doi.org/10.1007/s11222-015-9556-7 doi: 10.1007/s11222-015-9556-7
    [10] Y. Zhou, A. M. Johansen, J. A. Aston, Toward automatic model comparison: An adaptive sequential Monte Carlo approach, J. Comput. Graph. Statist., 25 (2016), 701–726. https://doi.org/10.1080/10618600.2015.1060885 doi: 10.1080/10618600.2015.1060885
    [11] R. M. Neal, Annealed importance sampling, Stat. Comput., 11 (2001), 125–139. https://doi.org/10.1023/A:1008923215028 doi: 10.1023/A:1008923215028
    [12] N. Chopin, A sequential particle filter method for static models, Biometrika, 89 (2002), 539–552. https://doi.org/10.1093/biomet/89.3.539 doi: 10.1093/biomet/89.3.539
    [13] P. Del Moral, A. Doucet, A.Jasra, Sequential monte carlo samplers, J. Royal Stat. Soc. Ser. B Stat. Methodol., 68 (2006), 411–436. https://doi.org/10.1111/j.1467-9868.2006.00553.x doi: 10.1111/j.1467-9868.2006.00553.x
    [14] F. Liang, W. H. Wong, Real-parameter evolutionary Monte Carlo with applications to Bayesian mixture models, J. Am. Stat. Assoc., 96 (2001), 653–666. https://doi.org/10.1198/016214501753168325 doi: 10.1198/016214501753168325
    [15] A. Jasra, D. A. Stephens, C. C. Holmes, On population-based simulation for static inference, Stat. Comput., 17 (2007), 263–279. https://doi.org/10.1007/s11222-007-9028-9 doi: 10.1007/s11222-007-9028-9
    [16] G. Evensen, The ensemble Kalman filter: Theoretical formulation and practical implementation, Ocean Dynam., 53 (2003), 343–364. https://doi.org/10.1007/s10236-003-0036-9 doi: 10.1007/s10236-003-0036-9
    [17] C. Schillings, A. M. Stuart, Analysis of the ensemble Kalman filter for inverse problems, SIAM J. Numer. Analys., 55 (2017), 1264–1290. https://doi.org/10.1137/16M105959X doi: 10.1137/16M105959X
    [18] D. Blömker, C. Schillings, P. Wacker, S. Weissmann, Well posedness and convergence analysis of the ensemble Kalman inversion, Inverse Probl., 35 (2019), 085004. https://doi.org/10.1088/1361-6420/ab149c doi: 10.1088/1361-6420/ab149c
    [19] F. J. Richards, A flexible growth function for empirical use, J. Exp. Bot., 10 (1959), 290–301. https://doi.org/10.1093/jxb/10.2.290 doi: 10.1093/jxb/10.2.290
    [20] M. Stoker, H. Rubin, Density dependent inhibition of cell growth in culture, Nature, 215 (1967), 171-–172. https://doi.org/10.1038/215171a0 doi: 10.1038/215171a0
    [21] G. Aubert, J. F. Aujol, A variational approach to removing multiplicative noise, SIAM J. Appl. Math., 68 (2008), 925–946. https://doi.org/10.1137/060671814 doi: 10.1137/060671814
    [22] Y. M. Huang, Yu-Mei, M. K. Ng, and Y. W. Wen, A new total variation method for multiplicative noise removal, SIAM J. Imag. Sci., 2 (2009), 20–40. https://doi.org/10.1137/080712593 doi: 10.1137/080712593
    [23] G. Steidl, T. Teuber, Removing multiplicative noise by Douglas-Rachford splitting methods, J. Math. Imag. Vis., 36 (2010), 168–184. https://doi.org/10.1007/s10851-009-0179-5 doi: 10.1007/s10851-009-0179-5
    [24] R. Šášik, E. Calvo, J. Corbeil, Statistical analysis of high-density oligonucleotide arrays: a multiplicative noise model, Bioinformatics, 18 (2002), 1633–1640. https://doi.org/10.1093/bioinformatics/18.12.1633 doi: 10.1093/bioinformatics/18.12.1633
    [25] A. M. Stuart, Inverse problems: A Bayesian perspective, Acta Numer., 19 (2010), 451–559. https://doi.org/10.1017/S0962492910000061 doi: 10.1017/S0962492910000061
    [26] N. Kantas, A. Beskos, A. Jasra, Sequential Monte Carlo methods for high-dimensional inverse problems: A case study for the Navier–Stokes equations, SIAM/ASA J. Uncertain. Quan., 2 (2014), 464–489. https://doi.org/10.1137/130930364 doi: 10.1137/130930364
    [27] P. Del Moral, Feynman-Kac Formulae, Springer-Verlag, New York, 2004. https://doi.org/10.1007/978-1-4684-9393-1
    [28] M. Bulté, J. Latz, E. Ullmann, A practical example for the non-linear Bayesian filtering of model parameters, in Quantification of Uncertainty: Improving Efficiency and Technology, Springer, Cham, (2020), 241–272. https://doi.org/10.1007/978-3-030-48721-8_11
    [29] H. Haario, E. Saksman, J. Tamminen, An adaptive Metropolis algorithm, Bernoulli, 7 (2001), 223–242.
    [30] Build models that drive breakthroughs, Technology Networks, 2018. Available from: https://www.technologynetworks.com/cancer-research/ebooks/build-models-that-drive-breakthroughs-311843.
    [31] K. Wrzesinski, A. Rogowska-Wrzesinska, R. Kanlaya, K. Borkowski, V. Schwämmle, J. Dai, et. al., The cultural divide: Exponential growth in classical 2D and metabolic equilibrium in 3D environments, PloS One, 9 (2014), e106973. https://doi.org/10.1371/journal.pone.0118050 doi: 10.1371/journal.pone.0118050
    [32] S. Ferson, W. L. Oberkampf, L. Ginzburg, Model validation and predictive capability for the thermal challenge problem, Computer Methods in Applied Mechanics and Engineering, 197 (2008), 2408–2430. https://doi.org/10.1016/j.cma.2004.04.030 doi: 10.1016/j.cma.2004.04.030
    [33] R. E. Kass, A. E. Raftery, Bayes factors, J. Am. Statist. Assoc., 90 (1995), 773–795. https://doi.org/10.1080/01621459.1995.10476572
    [34] O. J. Schmitz, Ecology and ecosystem conservation. Island Press, 2013.
    [35] A. Özkan, D. L. Stolley, E. N. Cressman, M. McMillin, S. DeMorrow, T. E. Yankeelov, et. al., Tumor microenvironment alters chemoresistance of hepatocellular carcinoma through CYP3A4 metabolic activity, Front. Oncol., 11 (2021). https://doi.org/10.3389/fonc.2021.662135
    [36] A. Ozkan, N. Ghousifam, P. J. Hoopes, T. E. Yankeelov, M. N. Rylander, In vitro vascularized liver and tumor tissue microenvironments on a chip for dynamic determination of nanoparticle transport and toxicity, Biotechnol. Bioeng., 116 (2019), 1201–1219. https://doi.org/10.1002/bit.26919 doi: 10.1002/bit.26919
    [37] A. Özkan, D. L. Stolley, E. N. Cressman, M. McMillin, T. E. Yankeelov, M. N. Rylander, CYP3A4 mediates chemoresistance controlled by cirrhosis and inflammation captured in vascularized hepatocellular carcinoma-on-a-chip, Submitted to Small, (2022).
    [38] P. Hartman, A lemma in the theory of structural stability of differential equations, Proceed. Am. Math. Soc., 11 (1960), 610–620. https://doi.org/10.2307/2034720 doi: 10.2307/2034720
    [39] D. M. Grobman, Homeomorphism of systems of differential equations (in Russian), Doklady Akademii Nauk SSSR, 128 (1959), 880–881.
  • Reader Comments
  • © 2022 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(2966) PDF downloads(134) Cited by(0)

Figures and Tables

Figures(10)  /  Tables(10)

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog