Research article

BCG and IL − 2 model for bladder cancer treatment with fast and slow dynamics based on SPVF method—stability analysis

  • Received: 21 November 2018 Accepted: 21 February 2019 Published: 10 June 2019
  • In this study, we apply the method of singularly perturbed vector field (SPVF) and its application to the problem of bladder cancer treatment that takes into account the combination of Bacillus Calmette–Guérin vaccine (BCG) and interleukin (IL)-2 immunotherapy (IL2). The model is presented with a hidden hierarchy of time scale of the dynamical variables of the system. By applying the SPVF, we transform the model to SPS (Singular Perturbed System) form with explicit hierarchy, i.e., slow and fast sub-systems. The decomposition of the model to fast and slow subsystems, first of all, reduces significantly the time computer calculations as well as the long and complex algebraic expressions when investigating the full model. In addition, this decomposition allows us to explore only the fast subsystem without losing important biological/ mathematical information of the original system.The main results of the paper were that we obtained explicit expressions of the equilibrium points of the model and investigated the stability of these points.

    Citation: OPhir Nave, Shlomo Hareli, Miriam Elbaz, Itzhak Hayim Iluz, Svetlana Bunimovich-Mendrazitsky. BCG and IL − 2 model for bladder cancer treatment with fast and slow dynamics based on SPVF method—stability analysis[J]. Mathematical Biosciences and Engineering, 2019, 16(5): 5346-5379. doi: 10.3934/mbe.2019267

    Related Papers:

    [1] K. E. Starkov, Svetlana Bunimovich-Mendrazitsky . Dynamical properties and tumor clearance conditions for a nine-dimensional model of bladder cancer immunotherapy. Mathematical Biosciences and Engineering, 2016, 13(5): 1059-1075. doi: 10.3934/mbe.2016030
    [2] Svetlana Bunimovich-Mendrazitsky, Yakov Goltser . Use of quasi-normal form to examine stability of tumor-free equilibrium in a mathematical model of bcg treatment of bladder cancer. Mathematical Biosciences and Engineering, 2011, 8(2): 529-547. doi: 10.3934/mbe.2011.8.529
    [3] Cameron Meaney, Gibin G Powathil, Ala Yaromina, Ludwig J Dubois, Philippe Lambin, Mohammad Kohandel . Role of hypoxia-activated prodrugs in combination with radiation therapy: An in silico approach. Mathematical Biosciences and Engineering, 2019, 16(6): 6257-6273. doi: 10.3934/mbe.2019312
    [4] Ben Sheller, Domenico D'Alessandro . Analysis of a cancer dormancy model and control of immuno-therapy. Mathematical Biosciences and Engineering, 2015, 12(5): 1037-1053. doi: 10.3934/mbe.2015.12.1037
    [5] Elzbieta Ratajczyk, Urszula Ledzewicz, Maciej Leszczynski, Avner Friedman . The role of TNF-α inhibitor in glioma virotherapy: A mathematical model. Mathematical Biosciences and Engineering, 2017, 14(1): 305-319. doi: 10.3934/mbe.2017020
    [6] Jason M. Graham, Bruce P. Ayati, Prem S. Ramakrishnan, James A. Martin . Towards a new spatial representation of bone remodeling. Mathematical Biosciences and Engineering, 2012, 9(2): 281-295. doi: 10.3934/mbe.2012.9.281
    [7] Silvia Martorano Raimundo, Hyun Mo Yang, Ezio Venturino . Theoretical assessment of the relative incidences of sensitive andresistant tuberculosis epidemic in presence of drug treatment. Mathematical Biosciences and Engineering, 2014, 11(4): 971-993. doi: 10.3934/mbe.2014.11.971
    [8] Adam Sullivan, Folashade Agusto, Sharon Bewick, Chunlei Su, Suzanne Lenhart, Xiaopeng Zhao . A mathematical model for within-host Toxoplasma gondii invasion dynamics. Mathematical Biosciences and Engineering, 2012, 9(3): 647-662. doi: 10.3934/mbe.2012.9.647
    [9] Joanna R. Wares, Joseph J. Crivelli, Chae-Ok Yun, Il-Kyu Choi, Jana L. Gevertz, Peter S. Kim . Treatment strategies for combining immunostimulatory oncolytic virus therapeutics with dendritic cell injections. Mathematical Biosciences and Engineering, 2015, 12(6): 1237-1256. doi: 10.3934/mbe.2015.12.1237
    [10] Gesham Magombedze, Winston Garira, Eddie Mwenje . Modelling the immunopathogenesis of HIV-1 infection and the effect of multidrug therapy: The role of fusion inhibitors in HAART. Mathematical Biosciences and Engineering, 2008, 5(3): 485-504. doi: 10.3934/mbe.2008.5.485
  • In this study, we apply the method of singularly perturbed vector field (SPVF) and its application to the problem of bladder cancer treatment that takes into account the combination of Bacillus Calmette–Guérin vaccine (BCG) and interleukin (IL)-2 immunotherapy (IL2). The model is presented with a hidden hierarchy of time scale of the dynamical variables of the system. By applying the SPVF, we transform the model to SPS (Singular Perturbed System) form with explicit hierarchy, i.e., slow and fast sub-systems. The decomposition of the model to fast and slow subsystems, first of all, reduces significantly the time computer calculations as well as the long and complex algebraic expressions when investigating the full model. In addition, this decomposition allows us to explore only the fast subsystem without losing important biological/ mathematical information of the original system.The main results of the paper were that we obtained explicit expressions of the equilibrium points of the model and investigated the stability of these points.


    Biological, chemical, physical and other phenomena with applications to various engineering science, are described by mathematical models that are described by a large set of complex linear or non-linear ordinary or partial differential equations. These systems of equations can be solved numerically. But one of the main drawbacks of a numerical solution of a system of equations is the loss of singular points of the system which can be critical for engineering applications. Another drawback in solving a complete set of equations is the time of computer running. Hence, there are different methods to reduce the system of the equations without losing information of the original one. These mathematical models have a number of essentially different time scales (i.e. rates of change) which correspond to sub-processes. These different time scales, in general, are not explicit for a given models, i.e., the hierarchy of the model is hidden. Once we expose the hierarchy, we can decompose the system of equations into fast and slow subsystems, and these decomposition allows one to apply different asymptotic methods. There are several asymptotic methods and numerical tools that can be applied to multi-scale systems. For example, the method of integral invariant manifold (MIM) [1,2,3,4], the iteration method of Fraser and Roussel [5,6,7,8], the computational singular perturbation (CSP) method [9,10], geometric singular perturbation theory [11,12,13], and the intrinsic low dimensional method (ILDM) which is a numerical method [14,15,16,17,18]. Each method has its advantages and disadvantages. The main disadvantage of all of these methods is that they are based on the fact that the hierarchy of the set of equations is given, i.e., all these model have the form of singularly perturbed system (SPS). According to what we have discussed above, we first need to expose the hierarchy of the given system. In this paper we introduced and applied a method called Singular Perturbed Vector Field (SPVF) [19,20] which is a new version of ILDM method. Since a large number of differential equations (a mathematical model) is presented in a hidden hierarchy form, no explicit time scale of the system is known. And hence, our aim is first to expose the time-scale of a given system, and then apply a reduction method. This is the main result of the SPVF method. The SPVF method transfers the original system to a form of singularly perturbed system (SPS), i.e., a system with an explicit hierarchy of the dynamical variables of the model. Once we transfer the system to SPS, the new system can be treated by the very powerful machinery of the standard SPS theory for model reduction and decomposition, as we have mentioned above without losing the essential dynamics of the original system.

    In addition, exposing the hierarchy of the model and presenting the system of equations in SPS, allow one to decompose the model into fast and slow subsystems and hence enables one to investigate the stability of the reduced model, which in itself is very important in models that describe cancer in general [21,22,23].

    In this research, we investigate the model of bladder cancer treatment that takes into account BCG and IL2 treatment presented in [24,25]. During bladder cancer therapy in the superficial phase the tumor is amenable to local excision, where small regions of cancerous tissue are surgically resected by direct inspection through the urethra, in a procedure called transurethral resection (TUR) [26]. To complete the procedure in the superficial stage, an adjuvant treatment is administered into the cavity is generally recommended to destroy any malignant cells that remain following the resection. The two complementary approaches to adjuvant treatment of superficial bladder cancer are the intravesical chemotherapy and the intravesical application of bacillus Calmette-Guerin (BCG) (immunotherapy).

    The mathematical model [24,25] describes the dynamics of BCG, APC, effector and tumor cells as a result of BCG and IL2 instillations. During this therapy, the dose of BCG and IL2 instills into the bladder via a catheter at every scheduled time. These instillations are described in the model via a set of Dirac delta functions. The disadvantage of using this function is that it is impossible to apply different asymptotic methods, except for solving the model numerically. Hence, we wrote an explicit function that describes the amount of BCG that is instilled into the bladder and which is depended on time. An explicit expression of function allowed us to overcome the disadvantages mentioned above as well as that we could control the dosages at different periods of BCG and IL2 pulsing.

    We investigated the mathematical model [24,25] that contains ten nonlinear ordinary differential equations by applying the SPVF method and exposing the hierarchy of the system explicitly. We used the eigenvectors of the system to change the coordinates and to present the model at SPS form with two fast equations and eighth slow equations. Revealing the hierarchy of the system allowed us to investigate the fast sub-system of the model, which reduces the complexity of the calculations and decreases the run-time of the computer. In addition, we presented the stability analysis of the reduced model. In general, knowledge about global fast manifolds is very important in the stability analysis of a reduced model described by a slow manifold. Now, depending on the investigated dynamical regime one can use the fast manifold as a manifold equation for the reduced space that can be represented by a low-dimensional manifold in the detailed linear vector space, which is given in an explicit form and proceed with the reduction procedure as projected to the manifold.

    In this section, we present the system of the governing equations (ODE) for bladder cancer treatment from [24,25], with BCG and IL2 combination therapy. The main assumptions of the model are as follows: A dose bk of BCG is instilled into the bladder every t days. It is given in 6-weekly intravesical instillations of 2108 c.f.u (colony-forming units, i.e., number of viable bacterial cells) [28]. After instillation, BCG accumulates close to the bladder wall. Upon binding to wall cells, BCG is internalized into the bladder and is processed by APCs [29] as well as the BCG binds and enters into malignant tumor cells [26]. The number of APCs increase as a result of recruitment in the response to bacterial infection. The bacterial infection stimulates the APCs to produce inflammatory cytokines such as IL2. Simultaneously, bacteria infect the occasional non-visualized residual cancer cell that remains after [30]. This causes the presentation of bacterial Ag on the tumor cell surface, which attracts APCs that ingest the entire host. Once a tumor cell has been ingested, tumor Ags are presented by the APCs. Due to the aforementioned inflammatory environment, created by the bacterial infection, APCs cause the CTLs to either mature and track bacteria Ag or mature and capture tumor cells according to their TAA [31]. This means that two CTL populations can destroy tumor cells either via the TAA mechanism (uninfected tumor cells) or via bacteria-associated Ag (infected tumor cells). The addition of exogenous IL2 is expected to create an inflammatory environment that stimulates the maturation of either CTL population and prolongs the life span of CTLs [32].

    We describe the instillation of BCG and IL2 into the bladder by a periodic function that at first time rises very rapidly (as exponentially). This actually describes the rapid instillation of the BCG and IL2 into the body, and then the function decreases in the form of decay exponential, which describes the removal of the BCG from the body. In order to define the function as described above in a periodic form, we used a variation of the Gamma probability density function in different intervals, then we summed all these functions.

    In the papers [24,25], they used the well known Dirac Delta function δ() to describe the instillation of a dose of BCG and IL2 into the bladder. Using this function, makes calculations very difficult and in some cases impossible to implement different asymptotic methods as we mentioned in the introduction. Hence, we proposed a new function (see Equation (2.11) bellow), instead of the Delta function, in order to overcome all the drawbacks listed above.

    Under the above assumptions and descriptions of the mechanism of the BCG-IL2 immune effect on the bladder the overall dynamics model and the interactions among ten different biological elements were presented the form:

    dBdt=J(t;α,β)p1ABp2BTuμBBFB(V), (2.1)
    dAdt=γ+ηABp1ABμAAθp3EBTiAFA(V), (2.2)
    dABdt=p1ABβABμA1ABFAB(V), (2.3)
    dATdt=θp3EBTiAλATuI2I2+gLβATμA1ATFAT(V), (2.4)
    dEBdt=βBABI2AB+g(p3Ti+μE1)EBFEB(V), (2.5)
    dETdt=βTATI2AT+g(p3Tu+μE2)ETFET(V), (2.6)
    dI2dt=(AB+AT+EB+ET)(q1q2I2I2+gL)+J(t;α,β)μI2I2FI2(V), (2.7)
    dTidt=p2BTup4EBTiFTi(V), (2.8)
    dTudt=rTu(1Tuk)p2BTu(λATTu+αETTuaT,βFβ+bT,βFβ+bT,β)I2gT(I2+gL)(Tu+gT)FTu(V), (2.9)
    dFβdt=aT,βTuμβFβFFβ(V), (2.10)

    where J(t)=J(t;α,β) is the sum of variation of Gamma Distributions functions with two free parameters α and β of the form

    J(t;α,β)=bktαΓ(α)βαmn=0eβ(t7n)α. (2.11)

    For more details about the Gamma Distributions functions please refer to the Appendix.

    Comment: At equation (2.1) we take into account a combination of different dose of BCG at a different period of BCG pulsing using the function J. At equation (2.7) we used with the same function J but only with one form of protocol: same dose of BCG at the same period of BCG pulsing that the BCG is instilled. From biological point of view, this means that we take into account different dosages of BGC at different times but IL2 is given to the patient at the same time when the BCG is given but during all the treatment the same dose of IL2 is given.

    The parameter bk indicates the dose BCG that instilled m times into the bladder inserted through the urethra every t days-period of BCG pulsing. In our analysis, we take into account different combinations of the amount of bk instilled into the bladder at different periods of BCG pulsing. Figure (1.1) present the function J that indicates given the same dose of BCG at the permanent period of BCG pulsing. Figure (1.2) present the function J that indicate a different dose of BCG at the same period of BCG pulsing. Figure (1.3) present the function J that indicates given a different dose of BCG at different periods of BCG pulsing. Figure (1.4) present the function J that indicates given the same dose of BCG at different periods of BCG pulsing.

    Figure 1.  The function described different/ same amount of bk instilled into the bladder at different/ same times.

    The initial conditions of the problem are

    B(0)=0,A(0)=104cells,AT(0)=AB(0)=EB(0)=ET(0)=0,I2(0)=50U/day,Ti(0)=0,Tu(0)=2106,Fβ(0)=0.02. (2.12)

    Comment: We assume that the tumor cells is an ellipsoid form with one radius of 1 mm [25], hence,

    Tu(0)=4π3r1r2,[Tu]=mm3. (2.13)

    Each ml contains 109 cells, and 1 ml is 1000 mm3. For example, lesion of tumor 13×15 mm in diameters, i.e. r1=6.5, r2=7.5, and hence Tu(0)=204 mm3 or contains about 0.204×109 cells.

    Equation (2.1) describes the dynamical rate of BCG level changes with time. This equation includes the function J that describes the combination between the different dose of BCG that is instilled into the bladder via a catheter inserted through the urethra at different periods of BCG pulsing. In addition this equation comprised of a positive term corresponding to BCG instillations, and of negative terms corresponding to the elimination of BCG by antigen-presenting cells (APCs) according to the rate coefficient p1, BCG tumor cell infection at a rate coefficient p2, and bacteria cell death with rate coefficient μB.

    Equation (2.2) is the dynamics of non-activated APCs, as described in B-M et al. [24], it is governed by two positive terms and three negative terms. The first positive term describes the normal influx of APCs to the tumor at a constant rate γ. The second positive term describes the recruitment of APCs due to bacterial infection at a rate coefficient η. The first negative term describes the activation of APCs by BCG at the rate coefficient p1. The second negative term is natural cell death at the rate coefficient μA. The last negative term accounts for the two-stages elimination of tumor cells, according to recent knowledge, first by effector CTL activity upon BCG-infected tumor cells, which leads to lysis of these cells and flooding of the tumor micro-environment with tumor antigens. The localized inflammatory response then attracts APCs, such as macrophages, which in turn eliminate uninfected tumor cells, according to the rate θp3.

    Equation (2.3) describes the dynamics of BCG-activated APCs. It is described by one positive term and two negative terms. The positive term is proportional to the number of non-activated APCs as well as BCG bacteria, with rate coefficient p1 (as in Equation (2.1)). The first negative term is the migration of the infected, activated APCs to the draining lymphoid tissues, at a rate of coefficient β1. The second negative term is the death of activated APCs at a rate of coefficient μA1.

    Equation (2.4) describes the tumor-Ag-activated APC (TAAAPC) dynamics. It is comprised of one positive term and three negative terms. The positive term describes the APCs which were activated by tumor antigen. The first negative term represents the tumor-Ag-activated APCs cells which destroy the uninfected tumor cells, with a rate coefficient λ. This term is multiplied by an IL2 dependent parameter with a saturation constant gL, to propose that in the absence of IL2, AT production ceases, while in the presence of external IL2, the production term is close to 1. The second negative term describes the migration of TAAAPC to the draining lymphoid tissues at a rate of coefficient β1. The third negative term denotes the natural death of TAAAPC at a rate coefficient μA1.

    Equation (2.5) describes the dynamics of effector CTLs that react to BCG infection. It is comprised of their migration rate, determined by their creation in the lymph nodes and subsequent migration to the bladder, inactivation rate, and their death rate. The migration element is proportional to AB and IL2, with a maximal rate of coefficient βB. This rate is brought to saturation by large numbers of AB, using a Michaelis-Menten saturation function, with Michaelis parameter g. The first negative term is the inactivation of effector CTLs via their encounter with infected tumor cells (Ti) at a success rate coefficient p3. The second negative term corresponds to the BCG-effector CTL (EB) cells' natural death rate μE1.

    Equation (2.6) describes the dynamics of effector cells reacting to tumor Ag. It is comprised of their migration rate, inactivation rate, and death rate. The migration element is proportional to AT and IL2 with a maximal rate coefficient βT. This rate is brought to saturation by large numbers of AT using a Michaelis-Menten saturation function, with Michaelis parameter g (as in Equation (2.5)). The first negative term describes the inactivation of effector CTLs via their encounter with uninfected tumor cells (Tu), at success rate coefficient p3. The second negative term describes the ET natural death rate, with a rate coefficient μE2.

    Equation (2.7) describes the IL2 dynamics. It is driven by a natural source, an external source, as well as sinking and degradation sources. The first two processes are positive and the last two are negative. They assume equal expression at the constant rate coefficient q1. This equation includes the function J as we described in Equation (2.1). I2 is consumed by APCs and CTLs. They assume that the rate of consumption is similar for both types of cells and denote its coefficient by q2. The consumption depends on I2 and is limited in a Michaelis-Menten fashion, with the Michaelis constant gL. They also introduce μI2, the I2 degradation rate coefficient.

    Equation (2.8) describes the dynamics of infected tumor cells depended on two mechanisms. The first corresponds only to the rate of bacterial infection of uninfected tumor cells, (Tu), according to rate coefficient p2. The second mechanism is the elimination of infected tumor cells (Ti) by their interaction with BCGCTL effector cells (represented by EB), at rate coefficient p4.

    Equation (2.9) describes the dynamics of uninfected tumor cells. It is comprised of three processes: one positive term, corresponding to natural tumor growth and two negative terms, corresponding to tumor infection by bacteria and tumor elimination by immune cells. The natural tumor growth is characterized by a maximal growth rate coefficient, r, which is limited by the maximal tumor cell number, k. The first negative term, due to bacterial infection, is characterized by a coefficient rate of p2. The second negative term is attributed both to the capture and elimination of Tu cells by APCs cells, which were activated by tumor-Ag at rate coefficient λ, and to the activity of TAACTL effectors, (ET), which destroy uninfected tumor cells, (Tu), at a rate coefficient α. The dependence in the equation of Tu on Fβ is decreasing from 1 to aT,β with Michaelis constant bT,β [33]. And then, there is a multiplication of those terms by an I2-dependent Michaelis Menten term, with Michaelis parameter gT, to propose that in the absence of I2, Tu cellular death does not occur. Since the tumor produces a variety of mechanisms in the biological settings that curtail the success of effector cell activity, they multiply I2/(I2+gL) by gT/(Tu+gT), to denote the inversely proportional reduction in effector cell activity rate, such that when Tu=0 the term is equal to 1 and when limTugT/(Tu+gT)=0. Note that although this factor can, in principle, nullify the efficacy of CTLs, this is not observed in cases of interest because Tuk [24].

    Equation (2.10) describes the dynamics of a transforming growth factor-beta (TGF-β, Fβ), as proportional to the tumor cell population, Tu, with aβ,T as a proportion coefficient and is destroyed at a rate of μβ proportional to Fβ.

    In this section, we present the effect of the function J on the model presented by [24].

    Figures (1) present the different combinations of BCG at different times.

    Figures (2)-(11) present the solution profiles of the model for different combinations of the BCG doses and periods of BCG pulsing (different/equal doses of BCG, different/ equal periods of BCG pulsing that the BCG introduced into the bladder).

    Figure 2.  The solutions profiles of B. 1 same amount of BCG at the permanent times, 2 different amount of BCG at the permanent times, 3 different amount of BCG at the different times, 4 same amount of BCG at the different times.
    Figure 3.  The solutions profiles of A. 1 same amount of BCG at the permanent times, 2 different amount of BCG at the permanent times, 3 different amount of BCG at the different times, 4 same amount of BCG at the different times.
    Figure 4.  The solutions profiles of AB. 1 same amount of BCG at the permanent times, 2 different amount of BCG at the permanent times, 3 different amount of BCG at the different times, 4 same amount of BCG at the different times.
    Figure 5.  The solutions profiles of AT. 1 same amount of BCG at the permanent times, 2 different amount of BCG at the permanent times, 3 different amount of BCG at the different times, 4 same amount of BCG at the different times.
    Figure 6.  The solutions profiles of EB. 1 same amount of BCG at the permanent times, 2 different amount of BCG at the permanent times, 3 different amount of BCG at the different times, 4 same amount of BCG at the different times.
    Figure 7.  The solutions profiles of ET.. 1 same amount of BCG at the permanent times, 2 different amount of BCG at the permanent times, 3 different amount of BCG at the different times, 4 same amount of BCG at the different times.
    Figure 8.  The solutions profiles of I2. 1 same amount of BCG at the permanent times, 2 different amount of BCG at the permanent times, 3 different amount of BCG at the different times, 4 same amount of BCG at the different times.
    Figure 9.  The solutions profiles of Ti. 1 same amount of BCG at the permanent times, 2 different amount of BCG at the permanent times, 3 different amount of BCG at the different times, 4 same amount of BCG at the different times.
    Figure 10.  The solutions profiles of Tu. 1 same amount of BCG at the permanent times, 2 different amount of BCG at the permanent times, 3 different amount of BCG at the different times, 4 same amount of BCG at the different times.
    Figure 11.  The solutions profiles of Fβ. 1 same amount of BCG at the permanent times, 2 different amount of BCG at the permanent times, 3 different amount of BCG at the different times, 4 same amount of BCG at the different times.

    Figure (2) presents the dynamical change of BCG in time. As we can see from the solution profiles of BCG for different combinations of BGC and time, the graphs behave in a wavy manner. This wavy behavior is due to the periodic function J(t) that takes into account different/ equal dose of BGC at different/ equal period of BCG pulsing. At the first stage, before the treatment, a patient is uninfected by BCG then B(0)=0. Then after the BCG instillation into the bladder and via a catheter inserted through the urethra. The BCG stimulates the body's immune system to destroy cancer cells, and then it leaves the body (by antigen-presenting cells (APCs), in addition BCG concentration decreases as a result of natural decay) in an exponential manner and repeat the process several times. Figure (2.1) refers to the solution profile of BCG with respect to the same amount at the same time. The BCG decreases to zero at the same rate of amount and time. Figure (2.2) refers to the solution profile of BCG with respect to different BCG but at the same time. At the beginning of treatment, the range of change of BCG is very large, and after that this range decreases to zero with respect to the same time. Figure (2.3) refers to the solution profile of BCG with different amounts at different time. According to this figure one can see that the quantities of BCG are change with time and reduced to zero according to the varying of the amount of BCG. Figure (2.4) refers to the solution profile of BCG with the same quantities of BCGat different time. It can be seen from this graph that indeed the BCG is reduced to zero but the amounts of the BCG are approximately the same with respect to time.

    Figure (3) presents the dynamics of non-activated APCs denoted by A. In all possible combinations of amount of BCG and time, we see that A, which starts from the initial condition 104, decreases very fast, but after that A increases and decreases periodically throughout the treatment and finally stabilized. The biological explanation for the reduction of the variable A from the initial condition is as follow: the activation of APCs by BCG that described by the term p1AB, and the capture of uninfected tumour cells (Tu) by APCs as well as the natural cell death of A, all these terms are very dominant compared to the sum of the term ηAB, which describes the recruitment of APC due to bacterial infection, and the parameter γ, we obtain that the dynamical variable is decreasing to its stable value.

    Figure (4) presents the dynamics of the variable AB the (APCs) cells that are activated by BCG. At the beginning of the process, there is a very large activity and interaction that behave exponentially and then a decrease of the activity as a result of the decline of the APC as seen from the graph (3). In addition, there is a decrease in activity as a result of migration of the infected and activated APCs to the draining lymphoid tissues, as well as the death of activated APCs, but this activity does not cause the variable to change. On the contrary, AB stabilizes at some constant value.

    Figure (5) presents the dynamics and interaction of Tumor Associated Antigen (TAA) with APCs cells. In general, T cells are activated by the link between TCR (T-Cell Receptor) and the tumor associated tumor (TAA) associated with the MHC (Major Histocompatibility Complex) which is presented by the Antigen (APC Presenting Cell), which is the initial "signal" of the T cell. The biological dynamics of this variable include a "competition" between the interaction of non-activated APCs, uninfected tumor cells and IL2 to the migration of the TAAAPC to the draining lymphoid tissues, and the natural death of TAAAPC. At first stage, the interaction TAAAPCIL2 overcomes to migration of the TAAAPC and the natural death of TAAAPC and after that, the process is reversed and AT decrease.

    Figure (6) presents the dynamical interaction of BCG and CTL effector cells. The biological dynamics of this variable EB is not like the normal dynamics of increase and decrease smoothly and continuously, but during the treatment, the interaction between BCG and CTL increases and decreases alternately until this interaction decreases. Since the main interaction depends on the migration rate that is determined by the creation in the lymph node and migration to the bladder, inactivation rate and their death rate. The term that causes to the variable EB to grow is an expression obeying MichaelisMenten saturated function's. This expression depends directly on AB, IL2 and MichaelisMenten constant g. The parameter g appears in the denominator of that expression. The numerator of this expression containing the rate coefficient βB108. This ratio between these two parameters, makes this expression very small relative to the (two) other terms of the equation that appear in it with a negative sign, what causes the derivative of EB to be negative and positive in certain points. As we will see later of the analysis of the results, the biological dynamics of IL2 rise and fall in a wave manner with a relatively small wavelength and very big amplitude. This amplitude is the main cause of the oscillations of the variable EB. The same behavior can be seen in AT's graph but in a much less prominent way. When we reduced the resolution of this graph we did see wavy behavior but with a very small wavelength and a very small amplitude. This is because of the interaction between BCG and CTL, where BCG have also a wavy behavior.

    The wavelength and amplitude of graphs (6.3) and (6.4) which refer to different doses of BCG at different periods of BCG pulsing and the same dose of BCG at permanent periods of BCG pulsing respectively, are more prominent compared to the other combinations of BCG and time, where the amplitude there and wavelength are smaller. In addition, the amplitude and wavelength of these graph are stabilized approximately to a certain value. In Figure (6.1) which refers to same dose of BCG at different periods of BCG pulsing, EB increases periodically very fast, then it is stabilized to a constant value, finally it decreases to zero very fast. The graph (6.2), which refers to the different dose of BCG at the same period of BCG pulsing, is on a rise in cycles and does not decrease. The explanation for this behavior is clear because during the treatment the interaction of BCG and CTL grows and goes down due to BCG's periodicity.

    Figure (7) presents the solution profile of the variable ET which describes the Tumor-effector CTLs interactions. The dynamics of ET are comprised of their migration rate, inactivation rate and death rate. The behavior of this variable is analogous to variables that periodically cycle through growth and decline. ETin Figures (7.2-4) behaves almost similarly. The different graph is the first one (7.24) which refers to same amount of BCG at the same time. The amplitude, as well as the wavelength of this graph, are smaller and denser than other graphs. Then drop fast rather than cyclic to zero as the others graphs.

    Figure (8) presents the dynamics of the variable interleukin (IL2). IL2 is a lymphokine that induces the proliferation of responsive T cells. In addition, it acts on some B cells, via receptor-specific binding as a growth factor and antibody production stimulant. The main aim of IL2 is the importance in stimulation of the adaptive immune system. In the equation of IL2 we insert the function J(t) that indicate different combination of BCG doses and periods of BCG pulsing. The natural sources of IL2 are the immune cells, i.e., the activated APCs and CTLs. In this work, an external source have been added to the treatment using the function J(t). In our numerical simulaiton we used with the same coefficient bk (which indicates the amount of external source) for both BCG and IL2. IL2 is injected into the bladder and modeled in the same manner as for BCG instillations. Graphs (7.1) and (7.4) behave the same (except the wavelength), a constant increase and decrease throughout the treatment until a final reduction at the end of the process. Graphs (7.2) and (7.3) also behave the same with the similar amplitude and wavelength. IL2 increases linearly during the treatment in periodically manner. The optimal solution for effective treatment is the growth of IL2 in controlled form. I.e., growth and stabilizing at a certain stage. As we can see, there is no graph solutions for an optimal behave of IL2. Because there are graphs that only rise and on the other hand there are graphs that only stabilize. Hence, by applying the SPVF method we obtain the optimal linear combinations of the dynamical variables of the system for an optimal treatment.

    Figure (9). Ti is the infected (or "marked" by the activation of APC (A), BCGAPC (AB), and BCGCTL (EB)) tumor cells. At the beginning of the BCG treatment, there are no infected tumor cells and hence Ti=0. After that, the number of marked cells increases because the cancer cells are marked by the others dynamics variables of the system. On the other hand, the cells Ti are eliminated due to the interaction with BCGCTL effector cells, and hence the graph of Ti decreases. The solutions profiles of (9.2)-(9.4) are optimal solutions since Ti cells decrease to zero. The solution of Ti, which refers to same amount of BCG at different times is not optimal, on the contrary, is worse for the patient. Because at some stage, although the tumor does not increase, it stabilizes, what can develop the disease and the recurrence of the tumor cells.

    Figure (10) is the solution of uninfected tumor cells. At the beginning of the treatment (t=0) the initial number of uninfected tumor cells in the urothelium before BCG treatment taken to be Tu(0)=2106. After that Tu decreases as expected to zero. The number of uninfected cells is decreasing and they become infected cells because of the interaction with the BCG, i.e., there is a transitions of cells from Tu to Ti by BCG action/ interaction. The growth of Tu is due to the natural tumor growth that proportional to T2u and Tu according to logistic tumor growth rate. According to [24], the number of Tu cells is limited by the tumor cell maximal number K=1011. Our simulation results show that Tu=2106<<k. The decrease of Tu is due to the tumor infection by bacteria as well as tumor elimination by immune system cells. The interactions and activations of the dynamical variables of the system such as APC (A), TAACTL effectors (ET), IL2 that dependent Michaelis-Menten term with a Michaelis parameter gL cause the decrease of Tu.

    Figure (11) describes the dynamics of a transforming growth factor-beta which behaves as the variable Tu, i.e., growth very fast and decrease to zero.

    General comment: The model's solutions in the new coordinates are irrelevant and biologically insignificant. What matters is the exposing of the system hierarchy and hence the decomposition of the model into fast and slow subsystems. And hence, the recipient of the linear combination of the variables of the (old) system with appropriate coefficients.

    As we can see from the above analysis of the results, one can not obtain the optimal treatment for the Bladder cancer (It is not known what combination of quantities with different or equal period of BCG and IL2 pulsing). Therefore, according to the SPVF method one can extract the optimal combination between the variables of the (old) system to achieve the maximal effect of the treatment and cause the elimination of the cancer cells.

    In this section we apply the Singular Perturbed Vector Field (SPVF) method to the bladder cancer model (2.1)–(2.10). The algorithm of the SPVF method can be found in paper [20].

    According to SPVF method the eigenvalues received by the algorithm of the SPVF

    λ1=6.3861019,λ2=8.1731016,λ3=985564.546,λ4=757899.292,λ5=665907.657,λ6=68785.561,λ7=5678.919,λ8=232.656,λ9=0.517,λ10=0.067. (3.1)

    According to the algorithm of the SPVF, the maximum gap is gapmaxi=λ2λ3=8.2921010. This means that the original system of equations can be decomposed into fast and slow subsystems, where the fast directions of the system are in the directions of the eigenvectors v1 and v2 that corresponds to the eigenvalues λ1 and λ2.

    The corresponds eigenvectors are

    v1=(0.4190.8050.4191.9591051.7581051.0341075.1021085.1531096.70510101.032109),v2=(0.5691.5921.5692.8801072.0251094.4511096.081087.59610115.0701086.377109),v3=(0.0173.6261060.0170.1560.1010.0040.0881.2381061.6091080.978),v4=(0.0045.0721070.0041.2081.7220.7450.57552.7261071.7581050.899), (3.2)
    v5=(0.0244.7071070.9400.7400.8970.6300.8775.4441042.61471050.182),v6=(0.97632.1341070.9570.8770.5620.8830.4010.0870.0870.213),v7=(0.1587.0921082.6830.0280.0984.01960.077.0050.0000.000),v8=(0.095.5251080.0360.0680.0550.9880.0150.0190.9540.335) (3.3)

    ,

    v9=(0.0014.74110100.0008.7871040.0020.9660.9060.98130.0463.736104),v10=(0.0047.83110120.0003.1891050.0070.7968.2751060.9440.4743.076107) (3.4)

    The above results show clearly the hierarchy of the given system in the new coordinates presentation. The fast direction of the system is in a parallel direction to the eigenvector v1. Then the less fast direction of the system is in a parallel direction to the eigenvector v2 and so on according to the descending order of eigenvalues since λ1>>λ2>>,...,>>λ10. The slow directions of the system are in parallel directions to the eigenvectors v3, v4, ..., v10 in descending order, where v10 is the slowest direction of the system.

    The next step of the SPVF method is transforming the model (2.1)–(2.10) using the above eigenvectors, hence using matrix in a form we write

    (x1x2x3x4x5x6x7x8x9x10)=(vt1vt2vt3vt4vt5vt6vt7vt8vt9vt10 )(BAABATEBETI2TiTuFβ) (3.5)

    where t refers to the transpose operation, i, e., it means that each eigenvector is placed in a row in the above matrix.

    For the convenience of the reader, we define the above system as matrix form hence, we denote the above matrix as A, the vectors of the variables of the original system as V=(B,A,AB,AT,EB,ET,I2,Ti,Tu,Fβ), and the vector of the new variables as U=(x1,x2,x3,x4,x5,x6,x7,x8,x9,x10). Hence, the system (3.5) can be rewritten as

    U=AV. (3.6)

    The next step is expressing the old variables of the system as a function of the new variables. For this aim, we multiply the set of equations (3.6) by the inverse matrix of A

    V=A1U. (3.7)

    In order to write the system of the ODE's model in the new coordinate, we take the derivative of the system (3.6) with respect to time, i.e.,

    dUdt=AdVdt, (3.8)

    and substitute the expressions of the RHS (right hand side) from the system (2.1)–(2.10) instead of dVdt in (3.8). In details

    dUdt=AdVdt=AF(V) (3.9)

    where

    F(V)=(FB(V)FA(V)FAB(V)FAT(V)FEB(V)FET(V)FI2(V)FTi(V)FTu(V)FFβ(V)). (3.10)

    is the RHS of the system (2.1)–(2.10). Substitute Equation (3.7) into Equation (3.9) and receive the model in the new coordinates with the initial conditions as

    dUdt=AF(A1U)G(U),U(0)=AV(0). (3.11)

    The model (2.1)–(2.10) that is rewritten in new coordinates as (3.11), is the model with the required hierarchy, and hence can be decomposed to fast and slow subsystems as follows

    fastsubsystem{ε1dx1dt=G1fast(x1,...,x10)ε2dx2dt=G2fast(x1,...,x10)
    slowsubsystem{dx3dt=G3slow(x1,...,x10)dx4dt=G4slow(x1,...,x10)dx5dt=G5slow(x1,...,x10)dx6dt=G6slow(x1,...,x10)dx7dt=G7slow(x1,...,x10)dx8dt=G8slow(x1,...,x10)dx9dt=G9slow(x1,...,x10)dx10dt=G10slow(x1,...,x10),

    where 0<ε1,ε2<<1, and the initial conditions are

    x1(0)=8050,x2(0)=15919.9,x3(0)=4.41532,x4(0)=63.9581,x5(0)=8.45245,x6(0)=173980.,x7(0)=3.49929,x8(0)=1.908106,x9(0)=92045.3,x10(0)=948000.

    The mathematical, and mainly biological significance of this decomposition is that it enables us to reduce the model and to investigate the dynamics of the system only on the fast subsystem without losing information if we would solve the complete system that includes all the variables of the system. The reduced model decreases the run-time of calculations and in some cases, one can even obtain an analytic expression for different variables of the reduced system.

    The next analysis is investigating the fast subsystem. We take into account only the variables that refer to the fast directions of the system i.e.,

    x1=0.419B+0.805A0.419AB+δ,x2=0.569B1.592A1.569AB+ξ, (3.12)

    where δ,ξ<<1. We take into account only three coordinates from the eigenvectors and neglect the rest because the other coordinates belong to the order of magnitude <107. The coordinates of the eigenvectors are actually the desire optimal combination, i.e., in the direction of v1 the dominant variable of the system is A (this variable will receive the largest weight) since its coefficient larger compared to the coefficients of B and AB. The other two variables will approximately the same weight since 0.41920.4192=1. The coefficients of the above linear combination indicate the proportions (relations) of the quantities (or percentages) that should be taken for the considered treatment for each variable of the old variables. The same analysis can be applied for the variable x2, which in the direction of the vector v2.

    In this section, we present a brief introduction to the fast and slow systems and investigate the stability of the considered model using the SPVF method.

    Once we apply the SPVF method, we can decompose the system into fast and slow subsystems. This means that the investigation of the model can be concentrated on the reduced model, i.e., one can investigate the stability analysis of the fast subsystem without loosing the information of the original system.

    Given a model that presented with a hidden hierarchy, by applying the SPVF method one can present the model with an explicit hierarchy, i.e., SPS form as

    εdvdt=Ffast(v,u,ε)dudt=Hslow(v,u,ε), (4.1)

    where 0<ε<<1, v=(v1,...,vmfast)Rmfast called fast variables which change fast compared to the slow ones u=(u1,...,ukslow)Rkslow and mfast+kslow=n (n is the dimension of the system. The functions F and H are assumed to be C of v, u and ε in (V×U)×I where V is an open subset of Rmfast, U is an open subset of Rkslow and I is an open interval containing ε=0.

    Setting τ=εt, the above model is transformed to

    dvdτ=Ffast(v,u,ε)dudτ=εHslow(v,u,ε). (4.2)

    Since 0<ε<<1 which describes the difference in time scales, than system (4.1) and (4.2) is called a slow and fast system, respectively.

    These systems are equivalent whenever ε0 and they are labeled singular perturbation problems when 0<ε<<1. The label "singular" stems in part from the discontinuous limiting behavior in the system (4.2) as ε0. At this case, when ε0, system (4.1) leads to differential-algebraic system

    0=Ffast(v,u,0)dudt=Hslow(v,u,0). (4.3)

    called reduced slow system which dimension decreases from mfast+kslow to kslow.

    On the other hand, system (4.2) reduces to an mfast-dimensional system called reduced fast system, with the variable u as a constant parameter:

    dvdτ=Ffast(v,u,0)dudτ=0. (4.4)

    By exploiting the decomposition into fast and slow reduced systems (4.4) and (4.3), the geometric approach reduced the full singularly perturbed system to separated lower-dimensional regular perturbation problems in the fast and slow regimes, respectively.

    In this paper we apply the SPVF method and present the considered model at the form of system (4.1).

    According to the manifolds theory approximation, the slow manifold is obtained by setting ε=0 in system (4.1)

    Mslow={Ffast(v,u,0)=0}, (4.5)

    In addition, we assumed that the solutions of (4.5) tend toward an equilibrium

    v=g(u), (4.6)

    which is the root of equation (4.5).

    The solutions of the system (4.1) have a fast transition from the initial points (or initial conditions) of the system to a point of the slow manifold Mslow. Hence, at the system (4.1) a slow motion takes place on the slow manifold, according to the equation

    dudt=Hslow(g(u),u). (4.7)

    This equation is called the reduced problem or the reduced model. This description of system (4.1) was given by Tikhonov [27].

    Follow the above theory of fast and slow system, the equilibrium points are obtained by solving the fast sub-system for the fast variables while setting all the slow variables as constant (the slow variables are "frozen" at their initial values and hence one can take u=u(0) as the constants);

    Ffast(v,uconst)=0. (4.8)

    Here, we denote with the star notation the equilibrium points which indicate the fast variables at their equilibrium values. Substitute the equilibrium points of the fast variables into the slow subsystem and solve this subsystem to find the equilibrium points of the slow variables, i.e.,

    Hslow(v,u)=0. (4.9)

    In order to analyze the stability of the above equilibrium points, one should examine the sign of the real part of the eigenvalues of the following Jacobian matrix at each equilibrium point of (4.8)–(4.9)

    (Ffast(v,u),Hslow(v,u))(v1,...,vmfast,u1,...,ukslow)|(v,u). (4.10)

    The eigenvalues with negative real parts are locally stable.

    Another method to find out the equilibrium points and their stability is to concentrate only on the fast subsystem, while the rest of the slow variables remaining constant. The advantage of this method is that we obtain more simple algebraic expressions and we reduce the running time of computer work. At this method, we solve only the fast subsystem (4.8) and the stability of the equilibrium points of the fast variables is given by analyzing the Jacobian matrix of the fast subsystem

    Ffast(v,uconst)(v1,...,vmfast)|(v,uconst) (4.11)

    which is a mfast×mfast matrix of partial derivatives of the fast subsystem with respect to the fast variables v while the slow variables are taken as constants.

    This means that the information about the stability of the model can be extracted from the analysis of the fast manifold.

    At the next section we apply the first method.

    In this section, we investigate the stability of the model that is the decomposed model into fast and slow subsystems.

    The following steps are implements for finding the fixed points in the new coordinates and determining their stability.

    1. substitute the slow variables as a constant into the fast subsystem (one can take the initial condition of the slow variables as the constants).

    2. setting all fast variables derivatives to zero i.e., solve the fast subsystem (with slow variables as constants) and find the equilibria points of the fast variables (x1,x2)

    G1fast(x1,x2,x3(0),...,x10(0))=0,G2fast(x1,x2,x3(0),...,x10(0))=0, (4.12)

    (here, we denoted the fast equilibrium points with star notation). Here, we have two equations with two unknown variables x1,x2.

    3. substitute the equilibrium points from step 2 (x1,x2) into the slow subsystem, solve the slow subsystem and find the equilibrium points of the slow variables (x3,...,x10)

    G3slow(x1,x2,x3,...,x10)=0,G4slow(x1,x2,x3,...,x10)=0,G5slow(x1,x2,x3,...,x10)=0,G6slow(x1,x2,x3,...,x10)=0,G7slow(x1,x2,x3,...,x10)=0,G8slow(x1,x2,x3,...,x10)=0,G9slow(x1,x2,x3,...,x10)=0,G10slow(x1,x2,x3,...,x10)=0. (4.13)

    Here, we have eight equations with eight unknown variables x3,...,x10.

    4. substitute the equilibrium points from steps 2 and 3 at the Jacobian matrix of the full system (the model at the new coordinates).

    5. compute the eigenvalues of the Jacobian matrix of the system (the model at the new coordinates) for each set of equilibrium point (the stable points are those with a negative real part of the eigenvalues).

    6. transform only the equilibrium points that are stable from steps 2 and 3 to the original coordinates using the inverse matrix of the eigenvectors, i.e., compute

    Vstable=A1Uistable (4.14)

    where i indicateS for different stable equilibrium points.

    It should be noted here that the equilibrium points we have received in the new coordinate system are functions of the parameters of the original model and a function of the initial conditions of the new model, i.e.,

    xi=xi(μA,μA1,...,gL;x1(0),...,x10(0)),i=1,...,10. (4.15)

    Comment 1: the initial condition as well as the equilibrium points {x1,x2,x3,...,x10} are biologically meaningless as there is no biological significance to the variables in the new coordinates. Hence, in order to understand the biological meaning of the above results, we should transform these points to the original coordinates using the inverse matrix of the eigenvectors A1. 2: Since the eigenvalues are invariant under change of coordinates, hence the stability equilibrium points will be remain stable equilibrium points at the original coordinates.

    By applying the above steps 16, we have found the following stability equilibrium point express in the original coordinates as a function of the parameters of the original model and the initial conditions at the new coordinates (For convenience, we denoted the initial condition xi(0) as x0i)

    B=J(t0;α,β)η+0.6224x01p1θgp2+0.3147x02p1μE1μB(μA11)0.4205x03(p2p1p3gLμB),A=0.6777x01ηγp3λη0.6548x02(p3θ)4.756x03(p1p3μA),AB=0.0976x01β(ηgLp1λ+g)+2.1732x03p1μA10.3524x04(βp1μA1)3.7689x05(μA1β)0.9994x06(βp11)η+0.6785x07(μA1βrp1gT)+0.5645x010p1,AT=0.5647x01(p3θ)g1L+0.7568x03(λ/μE2)+0.8765x04(βgLθ)ηp3+0.6544x05(1μA1gLp3)1.7666x06(βλ)0.5578x07(βgL)0.6574x010(μA1p3θ),EB=0.056x01μE1βg1T0.7658x03gp3ηλθ+0.9767x04(p3ηθ)+0.8678x05(gp3grkμB)/(gTaT,β)0.6755x07(μB)θbT,β+0.4565x08μE1p3gλ+0.7657x010(g/μI2p1),ET=2.8765x01gp3(μA1p3gTq2q1+gL)0.7655x03p3gL+0.5745x040.9876x05(μE2p3gq1(1+q2))+0.6534x06μE20.8765x07(p3)0.7547x08(p3gμE2)+0.8766x09g0.5436x010(p3gμE2μE2),I2=0.9878x01q1(q2)p3+0.6756x03(q2gL)p1p3θ+0.9282x04gLη+J(t0;α,β)0.7865x05(q1gLq2μBp1γμI2)0.7544x06(q1η)1(μI2gL)0.0098x07(q2gLq1q2)0.7865x08(gLq1)θγ,Ti=0.8767x01(p2)rgL+0.0092x03(p2)aT,βk0.5334x04(p21p4)0.0098x05p2p4(p1k)1+0.5434x06(1p4)+0.0195x07p4bT,βr0.7619x08(p2p4)+0.9454x09+0.2235x010,Tu=0.0661x01(rp2gLaT,β1)μBrgT+0.007x05λ(1α)0.6555x06(bT,βp2k)q1p1θ+0.2344x07(krλ)η0.4566x08(r)+0.6663x09(gTαp4),Fβ=0.5665x01(aT,βμBgT)μE2+0.0087x05(μB)p1ηγ+0.8655x06(μB1p4)rq1+1.7865x08r+0.4974x09λ bT,β (4.16)

    Here, we neglect elements of an order smaller than 104.

    As we have mentioned before, this equilibrium point is stable since the eigenvalues are invariant under change of coordinates.

    In order to complete writing the above stable equilibrium points in the original system, we write the initial conditions U as a function of the original initial conditions V using the matrix A as U(0)=AV(0) or in an explicit form

    x01=0.419B(0)+0.805A(0)0.419AB(0)1.959105AT(0)1.758105EB(0)+1.034107ET(0)+5.102108I2(0)5.153109Ti(0)+6.7051010Tu(0)+1.032109Fβ(0),x02=0.569B(0)1.592A(0)1.569AB(0)+2.880107AT(0)+2.025109EB(0)4.451109ET(0)+6.08108I2(0)7.5961011Ti(0)+5.070108Tu(0)6.377109Fβ(0),x03=0.017B(0)3.626106A(0)+0.017AB(0)0.156AT(0)0.101EB(0)0.004ET(0)+0.088I2(0)1.238106Ti(0)+1.609108Tu(0)+0.978Fβ(0),x04=0.004B(0)5.072107A(0)0.004AB(0)1.208AT(0)+1.722EB(0)0.745ET(0)0.5755I2(0)+2.726107Ti(0)1.758105Tu(0)0.899Fβ(0),x05=0.024B(0)4.707107A(0)+0.940AB(0)0.740AT(0)0.897EB(0)0.630ET(0)+0.877I2(0)+5.444104Ti+(0)2.6147105Tu(0)0.182Fβ(0),x06=0.9763B(0)+2.134107A(0)+0.957AB(0)+0.877105AT(0)0.562105EB(0)+0.883107ET(0)0.401108I2(0)0.087109Ti(0)+0.0871010Tu(0)+0.213109Fβ(0),x07=0.158B(0)+7.092107A(0)2.683AB(0)+0.028105AT(0)0.098105EB(0)4.0196107ET(0)0.07108I2(0)7.005109Ti(0)+0.0001010Tu(0)+0.000109Fβ(0),x08=0.09B(0)5.525108107A(0)0.036AB(0)0.068105AT(0)+0.055105EB(0)+0.988107ET(0)+0.015108I2(0)0.019109Ti(0)+0.9541010Tu(0)+0.335109Fβ(0),x09=0.001B(0)4.7411010A(0)+0.000AB8.787104AT+0.002EB(0)0.966ET(0)+0.906I2(0)0.9813Ti(0)+0.046Tu(0)3.736104Fβ(0),x010=0.004B(0)7.8311012A(0)0.000AB(0)+3.189105AT(0)0.007EB(0)0.796ET(0)8.275106I2(0)+0.944Ti(0)+0.474Tu(0)+3.076107Fβ(0). (4.17)

    By substitute equation (4.17) into equation (4.16), we receive the equilibrium point written completely in the original coordinates of the model. Our analysis for the equilibrium point includes different values of the function J(t0;α,β) i.e., different combinations of dosages and periods of BCG pulsing.

    Here, in table 14, we present the stable equilibrium point of the dynamical variable Tu for different values of bk at different periods of BCG pulsing (after 30 days, 44 days, 58 days and 65 days).

    Table 1.  Stable equilibrium points for a different combination of treatment. For the function J(1) which relates to the sub-graph 1 at Figure 1 that indicates the same dose of BCG at the permanent period of BCG pulsing.
    J(1)(t0;α,β) bk Tu
    t0=30 (days) 2108 2.032105
    t0=44 (days) 2108 5.024104
    t0=58 (days) 2108 2.786102
    t0=65 (days) 2108 1.000101

     | Show Table
    DownLoad: CSV
    Table 2.  Stable equilibrium points for a different combination of treatment. For the function J(2) which relates to the sub-graph 2 at Figure 1 that indicates a different doses of BCG at the permanent period of BCG pulsing.
    J(2)(t0;α,β) bk Tu
    t0=30 (days) 2.2107 4.009104
    t0=44 (days) 5.4107 1.745102
    t0=58 (days) 1.1108 0.078101
    t0=65 (days) 1.4108 0.034103

     | Show Table
    DownLoad: CSV
    Table 3.  Stable equilibrium points for a different combination of treatment. For the function J(3) which relate to the sub-graph 3 at Figure 1 that indicates a different dose of BCG at the different times.
    J(3)(t0;α,β) bk Tu
    t0=30 (days) 3.1107 1.756105
    t0=44 (days) 5.5107 2.659104
    t0=58 (days) 1.2108 0.876102
    t0=65 (days) 1.4108 0.001101

     | Show Table
    DownLoad: CSV
    Table 4.  Stable equilibrium points for a different combination of treatment. For the function J(4) which relate to the sub-graph 4 at Figure 1 that indicates the same dose of BCG at the different period of BCG pulsing.
    J(4)(t0;α,β) bk Tu
    t0=30 (days) 2108 5.773105
    t0=44 (days) 2108 3.067104
    t0=58 (days) 2108 2.965103
    t0=65 (days) 2108 0.060102

     | Show Table
    DownLoad: CSV

    Every table presents the stable equilibrium points expressed in the original variables of the bladder cancer model for different values of the Gamma functions, i.e., for different combinations of dosages and times. For example, 2 presents the stable equilibrium point for the function J(2) which relates to the sub-graph 2 at Figure 1 that indicates a different doses of BCG at the permanent period of BCG pulsing: after 30, 44 58 and 65 days. According to these results, we can see that this combination of treatment is optimal since it causes the variable Tu (cancer) to decrease to zero very fast compared to the other combinations of the function J, and this result is consistent with the results obtained from the numerical simulations presented in the graph of Tut.

    In this study, we improved the mathematical model for the treatment of bladder cancer using the probability density function in the form of a gamma function to describe the introduction of BCG and IL-2. Gamma function allows the delivery of treatment at different periods of pulsing and at different doses. The mathematical model includes ten nonlinear ordinary differential equations of first order. The model describes the dynamics of the combination between BCG and IL2 influence on the bladder tumor. The model describes the dynamics variability of the variables of the system involved in the interaction between the variables of the immune system and the variables of treatment in a way that the hierarchy is not explicit. We solved the model using numerical simulations for varying amounts of BCG and at varying period of BCG pulsing. Outcome following BCG therapy depends on the dose and period of pulsing [28].

    In order to obtain an optimal solution, we implemented the algorithm of the SPVF method. This method transforms the model, using eigenvectors, to a model presented in new coordinates in which the system can be splitted into a fast subsystem and a slow subsystem according to the hierarchy eigenvalues of the system. This hierarchy enables one to study the model using various asymptotic methods without losing any mathematical or biological information from the original system.

    After rewriting the model in the new coordinate using the eigenvectors received from the SPVF method, we exposed the hierarchy of the new model and received two fast variables and eight slow variables. This decomposition enabled us to apply the stability analysis to the fast subsystem. We found one stable equilibrium point express by the parameters of the original system and the initial conditions in the new coordinates. In order to obtain a biological meaning of these equilibrium point, we inversed and transformed the stable equilibrium points using the inverse matrix of the considered eigenvectors. And since the eigenvalues are invariant under change of coordinates, the stable equilibrium points remain stable at the original coordinates. After, applying this procedure, we received the equilibrium points expressed by the parameters of the original model and the initial conditions of the original model. The equilibrium points are also depend on the function J which takes into account different combination of dosages and periods of the pulsing therapy. We have found that the optimal combination of different dosages and periods of BCG pulsing at 30, 44, 58 and 65 days after the start of the treatment.

    Due to the SPVF method, we have explicitly received the equilibrium points, since this method enables us to reduce the dimension of the original model. We compared the results of the equilibrium point to the results we received from the model and found that they corresponded to our numerical simulations as one can see from the graphs.

    This work was supported by Jerusalem College of Technology.

    The authors declare that there is no Conflict of interest.

    Gamma distribution

    The general formula for the probability density function of the gamma distribution is:

    f(x)=(xμβ)α1exμββΓ(α), (5.1)

    for x μ, α,β>0, where α is the shape parameter, μ is the location parameter, β is the scale parameter, and Γ is the gamma function which has the formula

    Γ(α)=0tα1etdt, (5.2)

    with the following properties:

    1. Γ(1)=0exdx=1

    2. Γ(α+1)=αΓ(α)

    3. Γ(α)λα=0xα1eλxdx

    4. Γ(n)=(n1)!, n=1,2,...

    5. Γ(12)=π.

    The case where μ=0 and β=1 is called the standard gamma distribution and has the form of

    f(x)=xα1exΓ(α), (5.3)

    for x 0,α>0.

    Parameters

    μA=0.038 APC half life [days1]

    μA1=0.138 Activated APC half life [days1]

    μE1=0.19 Effector cells mortality rate W/O IL2 [days1]

    μE2=0.034 Effector cells mortality rate IL2 [days1]

    μB=0.1 BCG half life [days1]

    p1=1.25104 The rate of BCG binding with APC [cells1][days1]

    p2=0.028106 Infection rate of tumor cells by BCG [cells1][days1]

    p3=1.031010 Rate of E deactivation after binding with infected tumor cells [cells1][days1]

    p4=1.1106 Rate of destruction of infected tumor cells by effector cells [cells1][days1] λ=0.01106 Production rate of TAAAPC [days1]

    βB=1.45108 Recruitment rate of effector cells in response to signals released by BCG-infected and activated APC [cells1][days1][I12]

    βT=1.51106 Recruitment rate of effector cells in response to signals released by TAA-infected and activated APC [cells1][days1][I12]

    γ=4700 Initial APC cell numbers [cells1][days1]

    η=2.8106 Rate of recruited additional resting APCs [cells1][days1]

    r=0.00480.0085 Tumor growth rate [days1]

    β=0.034 Migration rate of TAAAPC and bacteria activation APC to the lymph node [cells1][days1]

    α=3.7106 Efficacy of an effector cell on tumor cells [cells1][days1]

    g = 1013 Michaelis-Menten constant for BCG activated CTLs and for TAACTLs [cells]

    gT=5.2103 Michaelis-Menten constant for tumor cells [cells]

    k=1011 Maximal tumor cell population [cells]

    q1=0.007 Rate of IL2 production [cells1][days1]

    q2=1.2103 The production of IL2 used for differentiation of effector cells IU production [cells1][days1]

    μI2=11.5 Degradation rate [days1]

    θ=0.01 Recruitment rate of Tumor-Ag-activated APC cells in response to signals released after binding effector cells that react to BCG infection, with infected tumor cells [1/cell1]

    aT,β=0.69 Michaelis-Menten saturation dynamics. The dependence of Fβ is decreasing from 0 to aT,β [dimensionless]

    bT,β=104 Michaelis constant [pg]

    μβ=166.32 The constant rate, account for degradation of Fβ [days1]

    gL=104 Michaelis-Menten constant for IL2 [cells]



    [1] V. M. Gol'dshtein and V. A. Sobolev, Singularity theory and some problems of functional analysis, Amer. Math. Soc., 1 (1992), 73–92.
    [2] V. I. Babushok and V. M. Gol'dshtein, Structure of the thermal explosion limit, Combust. Flame, 72 (1988), 221–226.
    [3] A. C. McIntosh, V. M. Gol'dshtein, I. Goldfarb, et al., Thermal explosion in a combustible gas containing fuel droplets, Combust. Th. Mod., 2 (1998), 153–165.
    [4] I. Goldfarb, V. M. Gol'dshtein, D. Katz, et al., Radiation effect on thermal explosion in a gas containing evaporating fuel droplets, Int. J. Ther. Sci., 46 (2007), 358–370.
    [5] M. R. Roussel and S. J. Fraser, Geometry of the steady-state approximation: perturbation and accelerated convergence methods, J. Chem. Phys., 93 (1990), 1072–1081.
    [6] M. R. Roussel and S. J. Fraser, Accurate steady-state approximations: implications for kinetics experiments and mechanism, J. Chem. Phys., 95 (1991), 8762–8770.
    [7] M. R. Roussel and S. J. Fraser, Global analysis of enzyme inhibition kinetics, textitJ. Chem. Phys., 97 (1993), 8316–8327.
    [8] M. R. Roussel and S. J. Fraser, Invariant manifold methods for metabolic model reduction, Chaos,196 (2001), 196–206.
    [9] A. Zagaris, H. G. Kaper and T. J. Kaper, Analysis of the computational singular perturbation reduction method for chemical kinetics, J. Non. Sci., 14 (2004), 59–91.
    [10] A. Zagaris, H. G. Kaper and T. J. Kaper, Fast and slow dynamics for the computational singular perturbation method, Soc. Indust. App. Math., 2 (2004), 613–638.
    [11] N. Berglunda and B. Gentzd, Geometric singular perturbation theory for stochastic differential equations, J. Diff. Eq., 191 (2003), 1–54.
    [12] N. Fenichel, Geometric singular perturbation theory for ordinary differential equations, J. Diff. Eq., 31 (1979), 53–98.
    [13] C. K. Jones, Geometric singular perturbation theory, 1609 of the series, Lec. Notes Math., Dyn. Syst. (2006), 44-118.
    [14] U. Maas and S. B. Pope, Implementation of simplified chemical kinetics based on intrinsic low-dimensional manifolds (PDF), Symposium (International) on Combustion, Twenty-Fourth Sym-posium on Combustion, (1992), 103–112.
    [15] U. Maas and S. B. Pope, Simplifying chemical kinetics: Intrinsic low-dimensional manifolds in composition space, Combust. Flame, 88 (1992), 239–264.
    [16] H. Bongers, J. A. Van Oijen and L. P. H. De Goey, Intrinsic low-dimensional manifold method extended with diffusion, Proc. Combust. Inst., 291 (2002), 1371–1378.
    [17] S. T. Alison, L. Whitehouse and L. Richard, The Estimation of Intrinsic Low Dimensional Man-ifold Dimension in Atmospheric Chemical Reaction Systems, Air Poll. Modell Simul., (2002), 245–263.
    [18] G. K. Hans and J. K. Tasso, Asymptotic analysis of two reduction methods for systems of chemical reactions, Phys. D, 165 (2002), 66–93.
    [19] V. Bykov, I. Goldfarb and V. Gol'dshtein, Singularly perturbed vector fields, J. Phys. Conf. Ser., 55 (2006), 28–44.
    [20] O. Nave, Singularly perturbed vector field method (SPVF) applied to combustion of monodisperse fuel spray, Diff. Eqs. Dyn. Syst., 27 (2018), 1–18.
    [21] M. Al-Tameemi, M. Chaplain and A. d'Onofrio, Evasion of tumours from the control of the im-mune system: consequences of brief encounters, Biol. Direct, 7 (2012), 1–22.
    [22] I. Kareva, F. Berezovskaya and C. Castillo-Chavez, Myeloid cells in tumour-immune interactions, J. Biol. Dyn., 4 (2010), 315–327.
    [23] L. Derbel, Analysis of a new model for tumor-immune system competition including long time scale effects, Math. Model Methods Appl. Sci., 14 (2004), 1657–1681.
    [24] K. E. Starkov and S. Bunimovich-Mendrazitsky, Dynamical properties and tumor clearance con-ditions for a nine-dimensional model of bladder cancer immunotherapy, Math. Biosci. Eng., 13 (2016), 1059–1075.
    [25] S. Bunimovich-Mendrazitsky, S. Halachmi and N. Kronik, Improving Bacillus Calmette-Guerin (BCG) immunotherapy for bladder cancer by adding interleukin 2 (IL−2): A mathematical model, Math. Med. Biol., 33 (2015), 1–30.
    [26] S. Brandau and H. Suttmann, Thirty years of BCG immunotherapy for non-muscle invasive blad-der cancer: a success story with room for improvement, Biomed. Pharmacother., 61 (2007), 299–305.
    [27] A. N. Tikhonov, Systems of differential equations containing small parameters multiplying the derivatives. Mat. Sborn., 31 (1952), 575–586.
    [28] R. Mahendran, Bacillus Calmette-Guerin immunotherapy-increasing dose as a means of improv-ing therapy?, Trans. Cancer Res., 6 (2017), 168–173.
    [29] A. Kiselyov, S. Bunimovich-Mendrazitsky and V. Startsev, Treatment of non-muscle invasive blad-der cancer with Bacillus Calmette-Guerin (BCG): biological markers and simulation studies, BBA Clin., 10 (2015), 27–34.
    [30] C. Pettenati and M. A.Ingersoll, Mechanisms of BCG immunotherapy and its outlook for bladder cancer, Nat. Rev. Urol., 15 (2018), 615–625.
    [31] A. H. Kitamur and T. Tsukamoto, Immunotherapy for urothelial carcinoma, Current status and perspectives, Cancers, 29 (2011), 3055–3072.
    [32] C. Yee, J. A. Thompson, D. Byrd, et al., Adoptive T cell therapy using antigen-specific CD8+ T cell clones for the treatment of patients with metastatic melanoma: in vivo persistence, migration, and antitumor effect of transferred T cells, Proc. Natl. Acad. Sci. U S A, 10 (2002), 16168–16173.
    [33] N. Kronik, Y. Kogan, P. G. Schlegel, et al., Improving T-cell immunotherapy for melanoma through a mathematically motivated strategy: efficacy in numbers?, J. Immunother., 35 (2012), 116–124.
  • This article has been cited by:

    1. Svetlana Bunimovich-Mendrazitsky, Leonid Shaikhet, Stability Analysis of Delayed Tumor-Antigen-ActivatedImmune Response in Combined BCG and IL-2Immunotherapy of Bladder Cancer, 2020, 8, 2227-9717, 1564, 10.3390/pr8121564
    2. OPhir Nave, Miriam Elbaz, Artificial immune system features added to breast cancer clinical data for machine learning (ML) applications, 2021, 202, 03032647, 104341, 10.1016/j.biosystems.2020.104341
    3. Joseph J. Barchi, Glycoconjugate Nanoparticle-Based Systems in Cancer Immunotherapy: Novel Designs and Recent Updates, 2022, 13, 1664-3224, 10.3389/fimmu.2022.852147
    4. Tiago Carvalho, Bruno Rodrigues Freitas, The local behavior around switching planes in a mathematical model to chemoimmunotherapy, 2023, 120, 10075704, 107186, 10.1016/j.cnsns.2023.107186
    5. T. Lazebnik, S. Yanetz, S. Bunimovich-Mendrazitsky, 2021, Chapter 9, 978-981-16-6296-6, 119, 10.1007/978-981-16-6297-3_9
    6. Ghassen Haddad, Amira Kebir, Nadia Raissi, Amira Bouhali, Slimane Ben Miled, Optimal control model of tumor treatment in the context of cancer stem cell, 2022, 19, 1551-0018, 4627, 10.3934/mbe.2022214
    7. Khaphetsi Joseph Mahasa, Rachid Ouifki, Amina Eladdadi, Lisette de Pillis, A combination therapy of oncolytic viruses and chimeric antigen receptor T cells: a mathematical model proof-of-concept, 2022, 19, 1551-0018, 4429, 10.3934/mbe.2022205
    8. Irina Volinsky, Svetlana Bunimovich-Mendrazitsky, Mathematical analysis of tumor-free equilibrium in BCG treatment with effective IL-2 infusion for bladder cancer model, 2022, 7, 2473-6988, 16388, 10.3934/math.2022896
    9. Sheng Zeng, Shaoqiang Xing, Yifei Zhang, Haifeng Wang, Qian Liu, Nano-Bacillus Calmette-Guérin immunotherapies for improved bladder cancer treatment, 2024, 25, 1673-1581, 557, 10.1631/jzus.B2300392
    10. Marom Yosef, Svetlana Bunimovich-Mendrazitsky, Mathematical model of MMC chemotherapy for non-invasive bladder cancer treatment, 2024, 14, 2234-943X, 10.3389/fonc.2024.1352065
    11. Yuko Shirono, Vladimir Bilim, Tsutomu Anraku, Hiroo Kuroki, Akira Kazama, Masaki Murata, Kaede Hiruma, Yoshihiko Tomita, Targeting Pro-Survival Autophagy Enhanced GSK-3β Inhibition-Induced Apoptosis and Retarded Proliferation in Bladder Cancer Cells, 2023, 30, 1718-7729, 5350, 10.3390/curroncol30060406
    12. Elizaveta Savchenko, Ariel Rosenfeld, Svetlana Bunimovich-Mendrazitsky, Mathematical modeling of BCG-based bladder cancer treatment using socio-demographics, 2023, 13, 2045-2322, 10.1038/s41598-023-45581-7
    13. Ming-Yuan Cao, Zhen-Dong Zhang, Xin-Rui Hou, Xiao-Ping Wang, The Potential Role of Non-coding RNAs in Regulating Ferroptosis in Cancer: Mechanisms and Application Prospects, 2024, 24, 18715206, 1182, 10.2174/0118715206322163240710112404
    14. Teddy Lazebnik, Svetlana Bunimovich-Mendrazitsky, Alex Kiselyov, Mathematical model for BCG-based treatment of type 1 diabetes, 2023, 622, 03784371, 128891, 10.1016/j.physa.2023.128891
  • Reader Comments
  • © 2019 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(5235) PDF downloads(650) Cited by(14)

Figures and Tables

Figures(11)  /  Tables(4)

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog