Loading [MathJax]/jax/output/SVG/jax.js
Research article Special Issues

Stochastic tumor-immune interaction model with external treatments and time delays: An optimal control problem


  • Herein, we discuss an optimal control problem (OC-P) of a stochastic delay differential model to describe the dynamics of tumor-immune interactions under stochastic white noises and external treatments. The required criteria for the existence of an ergodic stationary distribution and possible extinction of tumors are obtained through Lyapunov functional theory. A stochastic optimality system is developed to reduce tumor cells using some control variables. The study found that combining white noises and time delays greatly affected the dynamics of the tumor-immune interaction model. Based on numerical results, it can be shown which variables are optimal for controlling tumor growth and which controls are effective for reducing tumor growth. With some conditions, white noise reduces tumor cell growth in the optimality problem. Some numerical simulations are conducted to validate the main results.

    Citation: H. J. Alsakaji, F. A. Rihan, K. Udhayakumar, F. El Ktaibi. Stochastic tumor-immune interaction model with external treatments and time delays: An optimal control problem[J]. Mathematical Biosciences and Engineering, 2023, 20(11): 19270-19299. doi: 10.3934/mbe.2023852

    Related Papers:

    [1] Heping Ma, Hui Jian, Yu Shi . A sufficient maximum principle for backward stochastic systems with mixed delays. Mathematical Biosciences and Engineering, 2023, 20(12): 21211-21228. doi: 10.3934/mbe.2023938
    [2] Yuting Ding, Gaoyang Liu, Yong An . Stability and bifurcation analysis of a tumor-immune system with two delays and diffusion. Mathematical Biosciences and Engineering, 2022, 19(2): 1154-1173. doi: 10.3934/mbe.2022053
    [3] Haihua Zhou, Yaxin Liu, Zejia Wang, Huijuan Song . Linear stability for a free boundary problem modeling the growth of tumor cord with time delay. Mathematical Biosciences and Engineering, 2024, 21(2): 2344-2365. doi: 10.3934/mbe.2024103
    [4] Cristiana J. Silva, Helmut Maurer, Delfim F. M. Torres . Optimal control of a Tuberculosis model with state and control delays. Mathematical Biosciences and Engineering, 2017, 14(1): 321-337. doi: 10.3934/mbe.2017021
    [5] Bingshuo Wang, Wei Li, Junfeng Zhao, Natasa Trisovic . Longtime evolution and stationary response of a stochastic tumor-immune system with resting T cells. Mathematical Biosciences and Engineering, 2024, 21(2): 2813-2834. doi: 10.3934/mbe.2024125
    [6] 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
    [7] Giulio Caravagna, Alex Graudenzi, Alberto d’Onofrio . Distributed delays in a hybrid model of tumor-Immune system interplay. Mathematical Biosciences and Engineering, 2013, 10(1): 37-57. doi: 10.3934/mbe.2013.10.37
    [8] Donggu Lee, Sunju Oh, Sean Lawler, Yangjin Kim . Bistable dynamics of TAN-NK cells in tumor growth and control of radiotherapy-induced neutropenia in lung cancer treatment. Mathematical Biosciences and Engineering, 2025, 22(4): 744-809. doi: 10.3934/mbe.2025028
    [9] Fathalla A. Rihan, Hebatallah J. Alsakaji . Analysis of a stochastic HBV infection model with delayed immune response. Mathematical Biosciences and Engineering, 2021, 18(5): 5194-5220. doi: 10.3934/mbe.2021264
    [10] Yan Fu, Tian Lu, Meng Zhou, Dongwei Liu, Qihang Gan, Guowei Wang . Effect of color cross-correlated noise on the growth characteristics of tumor cells under immune surveillance. Mathematical Biosciences and Engineering, 2023, 20(12): 21626-21642. doi: 10.3934/mbe.2023957
  • Herein, we discuss an optimal control problem (OC-P) of a stochastic delay differential model to describe the dynamics of tumor-immune interactions under stochastic white noises and external treatments. The required criteria for the existence of an ergodic stationary distribution and possible extinction of tumors are obtained through Lyapunov functional theory. A stochastic optimality system is developed to reduce tumor cells using some control variables. The study found that combining white noises and time delays greatly affected the dynamics of the tumor-immune interaction model. Based on numerical results, it can be shown which variables are optimal for controlling tumor growth and which controls are effective for reducing tumor growth. With some conditions, white noise reduces tumor cell growth in the optimality problem. Some numerical simulations are conducted to validate the main results.



    Cancer is one of the most dangerous diseases in the world, and it affects both developing and developed nations. According to the World Health Organization, cancer is a prominent contributor to global mortality, with approximately 10 million deaths recorded in the year 2020 [1]. Through a multi-step or multifaceted process, cancer arises when normal cells turn into cancerous cells. It usually progresses from a pre-cancerous lesion to a fully cancerous cell. In this multistep mechanism, various genetic and molecular changes accumulate over time, ultimately leading to cancer. These transformations are the consequence of the interaction among a person's genetic circumstances and three classifications of external forces, such as physical, chemical and biological carcinogens [2]. An early diagnosis of cancer increases the likelihood of a favorable response to effective treatment, leading to a higher survival rate, reduced morbidity, and a more cost-effective treatment option. One of the most important and challenging questions is to understand cancer biology and how to treat it effectively. Therefore, many researchers have put time and efforts in creating and improving an efficient medication. Additionally, identifying the methods to support the patient improve their natural immunity that implement them to fight against tumor cells.

    Sophisticated mathematical models are necessary to accurately depict the intricate interplay between tumor cells and the immune system. Different mathematical models have been investigated and analyzed to recognize tumor-immune dynamics [3,4,5,6,7,8,9,10,11,12,13]. In [14], Kirschner and Panetta introduces the tumor-immune interaction model of three types of cell populations in vitro. The interaction between effector cells and tumor cells was mathematically modeled by Kuznetsov et al. [15] in one of the earliest proposals of such a model. An optimal control problem (OC-P) is typically formulated as a mathematical optimization problem, in which the objective is to find the control trajectory that minimizes or maximizes the cost function while satisfying any constraints on the state variables or control inputs of the system [16,17]. According to Rihan et al. [18] a study was carried out using an OC-P model for a delay differential model to examine tumor-immune interactions under immuno-chemotherapy.

    In actuality, deterministic models are not adequate to describe the dynamic process of the proliferation of cancer cells. Research has been conducted in this instance to amplify the deterministic models to stochastic counterpart [19,20,21,22,23,24]. Tumor-immune interactions are extremely complex, so it makes sense to include noise on the tumor-immune system in order to take into account a variety of relevant phenomena. An example of this may be variations in the intensity of neoantigens that stimulate the immune response, or changes in the expression of molecules that activate T cells; see [22,23,25]. As a result of the significant uncertainty inherent in the treatment process, authors in [26,27,28] modified the deterministic model to incorporate stochastic factors, such as variations in cellular reproduction and death rates, changes in the immune system's ability to fight tumor cells, and fluctuations in chemotherapy efficacy. In addition, the tumor-immune interaction model illustrates a delay between an immune response and a subsequent response by cancer cells [29]. A delay may affect the stability of the system, resulting in instability or bifurcation. In order to formulate effective therapies, it is crucial to understand how time-delays play a role in tumor-interaction models [29].

    Up to the best of authors' knowledge, few studies have been conducted so far in tumor-immune interactions with stochastic noises and time delays involving immunological boosts and chemotherapeutic treatment therapies. There is very little research relating stochastic models to interacting cells in cancer dynamics [30,31]. Drawing motivation from the aforementioned research studies, this paper attempts to investigate the impact of two time-delays and stochastic white noises on both the dynamics and optimal control performance of a model representing the interaction between tumors and the immune system. We provide a stochastic optimal control problem to reduce tumor growth and reduce the load tumors and increase the size of effector cells.

    This paper is structured as follows: Section 2 introduces a stochastic tumor-immune model with external treatments and time delay. Section 3 discusse the time-delayed stochastic model's global positive solution. The stationary distribution and extinction of tumor cells are discussed in Section 4. Section 5 examines the OC-P governed by the stochastic model using the stochastic maximal criterion. Section 6 presents numerical simulations incorporating various stochastic perturbations in order to validate the theoretical findings. Our conclusion is presented in Section 7.

    Here, we provide some preliminaries for the discussion that follows. Assume that (Ω,M,{M}t0,P) is a complete probability space with a filtration {Mt}t0 that meets the usual criteria. It is continuous with the right-hand side and M0 contains all P-null sets.

    Assume Z(t) is a regular time-homogeneous Markov process in C([τ,0];Rm+) and satisfies the following stochastic delay differential equation:

    dZ(t)=f(Z(t),Z(tτ),t)dt+g(Z(t),t)dW(t),fortτ,τ0, (1.1)

    with the initial value Z(s)=Z0C([τ,0];Rm+). W(t) stands for the m-dimensional standard Brownian motion defined on the complete probability space (Ω,M,{M}t0,P). Diffusion matrix Z(t) is then defined as follows:

    Π(Z)=(ςij(Z)),ςij(y)=gT(Z(t),t)g(Z(t),t).

    In order to have a nonnegative Lyapunov function V(Zt,t), it must be continuously twice differentiable in C2,1(C([τ,0];Rm+)×[τ,);R+) and once differentiable in t. Reference [21] defines the differentiable operator L of (1.1).

    Stochastic perturbations can influence the activation of immune cells, and it can affect the immune system's ability to detect and eliminate tumor cells. In addition, environmental fluctuations can impact the response to anti-cancer treatments, including immunotherapies. These perturbations can affect treatment efficacy, resistance development and long-term outcomes. First, we modify Kirschner and Panetta's model [14] which describes the dynamics of the activated effector cells E(t), tumor cells T(t) and Interleukin-2 (IL-2) cells I(t), with time delays τ1 and τ2. IL-2 plays a crucial role in regulating immune responses and promoting T cell proliferation and survival, it is also essential for generating effector and memory T cells. We consider the following assumptions: (ⅰ) The Lotka-Volterra form represents cell interaction. (ⅱ) The cancer cells are eradicated by the effector cells, while the IL-2 level increases as competition between tumor cells and effector cells increases. (ⅲ) Two distinct time-delays are incorporated, τ1 and τ2. The modified model takes the form

    dE(t)dt=αE(tτ1)T(tτ1)η1E(t)+μ1E(t)I(t)+e1,dT(t)dt=rT(t)(1βT(t))ξE(t)T(t),dI(t)dt=η2E(tτ2)T(tτ2)μ2I(t)+e2, (2.1)

    with the initial conditions E(θ)=σ1(θ),T(θ)=σ2(θ),I(θ)=σ3(θ), where σp(θ)0 for θ[τ,0),τ=max{τ1,τ2} and σp(0)>0,p=1,2,3. Tumor cell populations are reduced as a result of their interactions with effector cells E(t), which occur at distinct rates represented by ξE(t)T(t). By interacting with tumor effector cells ξE(t)T(t), immune effector cells E(t) reduce the population of tumor cells T(t) at a rate denoted by ξ. Effector cells are activated by IL-2 by stimulating their proliferation and differentiation, with rate μ1E(t)I(t). Effector cells are also stimulated by tumor cells, denoted by αE(tτ1)T(tτ1), where α represents the antigenicity rate of the tumor and τ1 is the duration over which IL-2 is induced in effector cells. η1E(t) represents the natural degradation of effector cells, while e1 represents the exogenous supply of effector cells. Proliferation and death of tumor cells are represented by the parameter r. β represents the biological environment's maximum carrying capacity for tumor cells T(t). η2E(tτ2)T(tτ2) represents the effector cells as the source of IL-2, which is induced by their interaction with tumor cells. Effectors and cancer cells compete at η2. The time delay between stimulating effector cells and tumor cells is represented by τ2. The parameter μ2 represents the degradation of IL-2 inherent to the system, whereas the variable e2 represents an exogenous IL-2 inflow. The parameters in model (2.1) are summarized in Table 1. See the Appendix, for the analysis of the deterministic model (2.1).

    Table 1.  Parameter definitions and values of model (2.1) (the unit is day1).
    Parameters Description Values Reference
    α the rate of antigenicity exhibited by the tumor 0.04 Assumed
    η1 death rate of effector cells 0.3743 [20]
    μ1 the level of cooperation between effector and IL-2 cells 0.035 [20]
    e1e2 the source of effector cells and IL-2 from external sources 0.1181 & 0.38 [20]
    r proliferation and mortality rate of tumor cells 0.8636 Assumed
    β1 the biological environment's tumor-carrying capacity 0.002 [20]
    ξ inactivation rate of tumor cells by effector cells 1 [20]
    η2 the rate of competition between effector 0.01 Assumed
    and cancer cells
    μ2 the rate of loss of IL-2 cells. 0.055 [20]

     | Show Table
    DownLoad: CSV

    Most biological phenomena are characterized by random fluctuations, particularly variations in the intensity of neoantigens that trigger immune responses or the expression of molecules that activate T cells. Due to the complexity of the interaction between tumor cells and immune effectors, noise related to the tumor-immune system is justified. Since tumor-immune cells may have small populations, the impact of random fluctuations is significant. To investigate the impact of stochastic fluctuations within a deterministic model, there are primarily two methodologies [32,33,34,35]. One or more relevant parameters can be substituted with their stochastic counterparts in a deterministic model. In this process, white or colored noises are introduced to the deterministic parameters to introduce stochastic perturbations. Alternatively, it is possible to introduce a stochastic driving force directly into the deterministic model instead of modifying specific parameters. In the current study, we adopt the second alternative approach, wherein we assume the occurrence of stochastic perturbations in the variables, throughout the boundary and the interior equilibrium point. The perturbations are assumed to follow a white noise distribution, whose magnitude is proportional to the distances between equilibrium values E(t), T(t), and I(t). The model structure is as follows:

    dE(t)=[αE(tτ1)T(tτ1)η1E(t)+μ1E(t)I(t)+e1]dt+ν1H1dW1,dT(t)=[rT(1βT)ξE(t)T(t)]dt+ν2H2dW2,dI(t)=[η2E(tτ2)T(tτ2)μ2I(t)+e2]dt+ν3H3dW3, (2.2)

    with initial conditions

    E(θ)=σ1(θ)0;T(θ)=σ2(θ)0;I(θ)=σ3(θ)0,θ[τ,0),σp(0)>0,p=1,2,3. (2.3)

    Wi(i=1,2,3) stand for the independent Brownian motions that are stated on a complete probability space (Ω,M,{M}t0,P) with a filtration {Mt}t0 satisfying the usual conditions and M0 contains all P-null sets. Let R+=(0,+),Rm+={(y1,y2,,ym)Rm|yp>0,p=1,2,,m}, Hi=Hi(E,I,T), (i=1,2,3) are locally Lipschitz-continuous functions, such that Hi has the following forms:

    ● In the first form R1: H1=E, H2=T, and H3=I, i.e., the environmental impact on the cell described by stochastic perturbation [36].

    ● In the second form R2: H1=EE0, H2=TT0, and H3=II0, where S=(E0,T0,I0) is the equilibrium point of (2.1), which investigates the behavior of stochastic perturbations around the equilibrium point [37].

    When examining the dynamics of the proposed system, the positivity of the solution is of primary importance. The following section provides a detailed analysis of the solution's positivity.

    Here, we examine the global positivity criteria pertaining to the model (2.2) of the interaction between tumor and immune cells. In order to determine the positivity of the solutions of model (2.2), it is necessary to consider the Banach space C=C([τ,0],R3+), which has continuous functions that map the interval [τ,0] into R3+ with topological uniform convergence, where, R3+={(E0,T0,I0)|E00;T00;I00}. The following two theorems demonstrate that the stochastic model can have a positive global solution (2.2) with forms R1 and R2.

    Theorem 3.1. For any given initial condition (E0,T0,I0)R3+, the model (2.2) with form R1 for all tτ, where τ=max{τ1,τ2}, has a solution which is almost surely unique and positive.

    Proof. For any initial condition (E0,T0,I0)R3+, as the coefficients of system (2.2) satisfy the local Lipschitz condition, so system (2.2) has a unique local solution (E(t),T(t),I(t))ont[τ,τe) almost surely, where τe stands for the explosion time. The target is to show that this solution is global i.e. τe= with probability one. Assume c01 to be sufficiently large such that E(θ),T(θ)I(θ)(θ[τ,0]) are lying in the interval [1c0,c0]. For each cc0,cN, define the stopping time

    τc=inf{t[τ,τe):min{E(t),T(t),I(t)}1cormax{E(t),T(t),I(t)}c}.

    Assume infϕ=. Therefore, τc is increasing as c. Let τ=limcτc, then ττe. One needs to show that τ= with probability one, then τe= a.s. and (E(t),T(t),I(t))R3+ for all tτ. When it is incorrect, there is a pair ι(0,1) and ˜T>0 such that P{τ˜T}>ι. Hence, there is an integer c1c0 such that

    P{τc˜T}ι,cc1. (3.1)

    By employing the Lyapunov functional methodology, it can be observed that the system represented by Eq (2.2) exhibits the presence of a positive solution that is global in nature. Thus, we define a C2-function G(E,T,I):R3+R+ by

    G(E,T,I)=(E1lnE)+(α+η2ξ)(T1lnT)+(I1lnI)+αttτ1E(s)T(s)ds+η2ttτ2E(s)T(s)ds.

    Applying Itˆo's formula to G(E,T,I), we get

    dG(E,T,I)=LG(E,T,I)dt+ν1(E1)dW1+ν2(α+η2ξ)(T1)dW2+ν3(I1)dW3, (3.2)

    where

    LG(E,T,I)=αETη1E+μ1EI+e1αE(tτ1)T(tτ1)E+η1μ1Ie1E+(α+η2ξ)rT(1βT)(α+η2ξ)ξET(α+η2ξ)r(1βT)+(α+η2ξ)ξE+η2ETμ2I+e2η2E(tτ2)T(tτ2)I+μ2e2I+ν1+ν2+ν32(α+η2(α+η2ξ)ξ)ET+μ1EI+((α+η2ξ)ξη1)E+(α+η2ξ)r(1+β)T(μ1+μ2)I+e1+e2+η1+μ2r+ν1+(α+η2ξ)ν2+ν32supER+{((α+η2ξ)ξη1)E+μ1E2}+supIR+{μ1I2(μ1+μ2)I}+supTR+{(α+η2ξ)r(1+β)T(α+η2ξ)rβT2}+e1+e2+η1+μ2r+ν1+(α+η2ξ)ν2+ν32N, (3.3)

    and N>0 is a constant which is independent of E(t),T(t),I(t). Thus,

    dG(E,T,I)Ndt+ν1(E1)dW1+(α+η2ξ)ν2(T1)dW2+ν3(I1)dW3. (3.4)

    Integrating (3.4) from 0 to τc˜T=min{τc,˜T} and then taking the expectation E on both sides, we get

    EG(E(τc˜T),T(τc˜T),I(τc˜T))EG(E(0),T(0),I(0))+N˜T. (3.5)

    Let Ωc={τc˜T}, for cc1 and in view of (3.1), we obtain P(Ωc)ι such that, for every ωΩc, there is at least one of E(τc,ω),T(τc,ω), or I(τc,ω) equaling either c or 1c, and then one obtains

    G(E(τc˜T),T(τc˜T),I(τc˜T))(c1lnc)(1c1ln1c). (3.6)

    In view of (3.5), we have

    EG(E(0),T(0),I(0))+N˜TE[1Ωn(ω)G(E(τc,ω),T(τc,ω),I(τc,ω))]ι(c1lnc)(1c1ln1c), (3.7)

    where 1Ωc stands for the indicator function of Ωc. As c, we have >EG(E(0),T(0),I(0))+N˜T=, which leads to a contradiction. Thus, it can be concluded that τ= with probability one, which proves the theorem.

    Theorem 3.2. There exists a unique solution (E(t),T(t),I(t)) of tumor-immune model (2.2) with form R2, for all tτ, where τ=max{τ1,τ2}, and any initial condition (E0,T0,I0)R3+. The solution will remain in R3+, that is (E(t),T(t),I(t))R3+ for all tτ almost surely.

    Proof. Herein, we establish a C2-function U(E,T,I):R3+R+ as follows:

    U(E,T,I)=(E1lnE)+(α+η2ξ)(T1lnT)+(I1lnI)+αttτ1E(s)T(s)ds+η2ttτ2E(s)T(s)ds.

    Applying Itˆo's formula to U(E,T,I), one obtains

    dU(E,T,I)=[(11E)[αE(tτ1)T(tτ1)η1E+μ1EI+e1]+ν212(1E0E)2+(11T)[rT(1βT)ξET]+ν222(1T0T)2+(11I)[η2E(tτ2)T(tτ2)μ2I+e2]+ν232(1I0I)]dt+ν1(11E)(EE0)dW1(t)+ν2(11T)(TT0)dW2+ν3(11I)(II0)dW3(t). (3.8)

    Therefore, LU(E,T,I):R3+R+ is as follows:

    LU(E,T,I)supER+{((α+η2ξ)ξη1)E+μ1E2}+supIR+{μ1I2(μ1+μ2)I}+supTR+{(α+η2ξ)r(1+β)T(α+η2ξ)rβT2}+e1+e2+η1+μ2r+ν1+(α+η2ξ)ν2+ν32+ν212(1E0E)2+ν222(1T0T)2+ν232(1I0I)2ˆN. (3.9)

    The remainder of the proof closely resembles the proof of Theorem 3.1.

    The objective of this section is to establish adequate criteria for the presence of a unique ergodic stationary distribution.

    For each t0, and probability measure μ on (C([τ,0];Rm+),M[τ,0]), where M[τ,0] is the associated Borel σ-algebra in [τ,0], consider the probability measure μPt on (C([τ,0];Rm+),M[τ,0]) defined by

    (μPt)(Δ)=C([τ,0];Rm+)Pt(y,Δ)μ(dy),forΔM[τ,0].

    Definition 4.1. (Stationary Distribution [38]). A stationary distribution for (1.1) is a probability measure π on (C([τ,0];Rm+),M[τ,0]) such that (πPt)(Δ)=π(Δ) for all t0 and ΔM[τ,0].

    In order to analyze the existence of the stationary distribution of the SDDE system (2.2), it is enough to take m=3. For the stochastic model (2.2), the threshold parameter is defined as follows:

    ˆT0=αe1e2+r^η1^μ2^η1^μ2(ν22/2), (4.1)

    such that ^η1=η1+ν212, ^μ2=μ2+ν232.

    Theorem 4.1. The tumor-immune model (2.2) admits a stationary distribution π(.) if ˆT0>1 for any initial conditions (2.3).

    Proof. According to Theorem 3.1, it has been established that there is a unique solution (E(t),T(t),I(t))R3+ on tτ for (2.2) for any given initial values (2.3). Thus,

    Step 1: The diffusion matrix of stochastic delayed tumour immune model (2.2) is

    A(E,T,I)=(ν21E2000ν22T2000ν23I2). (4.2)

    Let X be any bounded domain in R3+, then there exists a positive constant d0=min{ν21E2,ν22T2,ν23I2,(E,T,I)˜X} such that 3i,j=1ςij(y)aiaj=ν21E2a21+ν22T2a22+ν23I2a23d0|a|2 for all (E,T,I)˜X,aR3+. This implies that the smallest eigenvalue of the diffusion matrix Π(E,T,I) is bounded away from zero.

    Step 2: We construct a nonnegative twice continuously differentiable function F:R3+R is introduced as follows:

    F(E,T,I)=Q(lnTc1lnEc2lnI)lnE+lnTIlnI=QF1+F2, (4.3)

    where c1=e1e2α^η12^μ2 and c2=e1e2α^η1^μ22.

    Select a constant Q>0 large enough such that

    QΩ+ρ2, (4.4)

    where Ω=(r+αe1e2^η1^μ2)ν222>0 since Ts0>1, and ρ=max{ρ1,ρ2,ρ3},

    ρ1=sup(E,T,I)R3+{QξE+η1μ1I+rrβTξE+η2E(tτ2)T(tτ2)μ2I+e2+ν21+ν222η2E(tτ2)T(tτ2)I+μ2+ν232e2I}<,ρ2=sup(E,T,I)R3+{QrβT+η1μ1I+rrβTξE+η2E(tτ2)T(tτ2)μ2I+e2+ν21+ν222η2E(tτ2)T(tτ2)I+μ2+ν232e2I}<,ρ3=sup(E,T,I)R3+{QξE+QrβT+η1μ1I+rrβTξE+η2E(tτ2)T(tτ2)μ2I+e2+ν21+ν222η2E(tτ2)T(tτ2)I+μ2+ν232e2I}<. (4.5)

    By the application of Itˆo's formula to F1, the resulting expression is obtained as:

    LF1=r(1βT)+ξE+ν222c1αE(tτ1)T(tτ1)E+c1(η1+ν212)c1μ1Ic1e1Ec2η2E(tτ2)T(tτ2)I+c2(μ2+ν232)c2e2I,33αe1e2c1c2+c1(η1+ν212)+c2(μ2+ν232)r+rβT+ξE+ν222,ν222(r+αe1e2^η1^μ2)+rβT+ξE,:=Ω+rβT+ξE. (4.6)

    In the same manner, we can get

    LF2=αE(tτ1)T(tτ1)E+η1μ1Ie1E+rrβTξE+η2E(tτ2)T(tτ2)μ2I+e2+ν21+ν222ν232η2E(tτ2)T(tτ2)I+μ2+ν232e2I. (4.7)

    From Eqs (4.6) and (4.7), one gets

    L˜FQ(Ω+rβT+ξE)αE(tτ1)T(tτ1)E+η1μ1Ie1E+rrβTξE+η2E(tτ2)T(tτ2)μ2I+e2+ν21+ν222ν232η2E(tτ2)T(tτ2)I+μ2+ν232e2I,Q(Ω+rβT+ξE)+η1μ1I+rrβTξE+η2E(tτ2)T(tτ2)μ2I+e2+ν21+ν222η2E(tτ2)T(tτ2)I+μ2+ν232e2I. (4.8)

    Let us define a bounded closed set for any value of ϵ>0

    Z={(E,T,I)R3+:ϵE1ϵ,ϵT1ϵ,ϵ3I1ϵ3}. (4.9)

    To enhance intuitiveness, we divide R3+Z=6i=1Zi, into the following six regions:

    Z1={(E,T,I)R3+;0<E<ϵ},Z2={(E,T,I)R3+;0<T<ϵ},Z3={(E,T,I)R3+;0<I<ϵ3},Z4={(E,T,I)R3+;E>1ϵ},Z5={(E,T,I)R3+;T>1ϵ},Z6={(E,T,I)R3+;E>ϵ ,T>ϵ,I>1ϵ3}. (4.10)

    To prove L˜F1 for any (E,T,I)R3+Z=6i=1Zi, we consider six cases as follows:

    C.Ⅰ: For any (E,T,I)Z1

    L˜FQ(Ω+rβT+ξE)+η1μ1I+rrβTξE+η2E(tτ2)T(tτ2)μ2I+e2+ν21+ν222η2E(tτ2)T(tτ2)I+μ2+ν232e2IQΩ+QrβT+ρ1QΩ+Qrβϵ+ρ1, (4.11)

    from Eq (4.4) and choosing ϵ1Qrβ.

    C.Ⅱ. For any (E,T,I)Z2

    L˜FQ(Ω+rβT+ξE)+η1μ1I+rrβTξE+η2E(tτ2)T(tτ2)μ2I+e2+ν21+ν222η2E(tτ2)T(tτ2)I+μ2+ν232e2IQΩ+QξE+ρ2QΩ+Qξϵ+ρ1 (4.12)

    from Eq (4.4) and choosing ϵ1Qξ.

    C.Ⅲ. For any (E,T,I)Z3

    L˜FQ(Ω+rβT+ξE)+η1μ1I+rrβTξE+η2E(tτ2)T(tτ2)μ2I+e2+ν21+ν222η2E(tτ2)T(tτ2)I+μ2+ν232e2Iη2E(tτ2)T(tτ2)I+ρ3η2ϵ2ϵ3+ρ31 (4.13)

    by choosing ϵη2ρ3.

    C.Ⅳ. For any (E,T,I)Z4

    L˜FQ(Ω+rβT+ξE)+η1μ1I+rrβTξE+η2E(tτ2)T(tτ2)μ2I+e2+ν21+ν222η2E(tτ2)T(tτ2)I+μ2+ν232e2IξE+ρ3ξϵ+ρ31 (4.14)

    by choosing ϵξρ3.

    C.Ⅴ. For any (E,T,I)Z5

    L˜FQ(Ω+rβT+ξE)+η1μ1I+rrβTξE+η2E(tτ2)T(tτ2)μ2I+e2+ν21+ν222η2E(tτ2)T(tτ2)I+μ2+ν232e2IrβT+ρ3rβϵ+ρ31 (4.15)

    by choosing ϵrβρ3.

    C.Ⅵ. For any (E,T,I)Z6

    L˜FQ(Ω+rβT+ξE)+η1μ1I+rrβTξE+η2E(tτ2)T(tτ2)μ2I+e2+ν21+ν222η2E(tτ2)T(tτ2)I+μ2+ν232e2Ir(μ1+μ2)I+ρ3(μ1+μ2)ϵ3+ρ31 (4.16)

    by choosing ϵ3(μ1+μ2)ρ3.

    This implies that if the solution Z(t)=(E(t),T(t),I(t))R3+/Z of the stochastic system (2.2), the mean times τ1τ2 at which the path issuing from Z(t) reaches the set Z is finite for every compact subset XR3+.

    Definition 4.2. [21] If limtT(t)=0 a.s, the tumor T(t) is said to extinct with probability one.

    Theorem 4.2. Suppose that (E(t),T(t),I(t)) is the solution set of the tumor-immune interaction model system (2.2) along with the initial condition (E0,T0,I0)R3+. If r<ν222, then

    limtsuplnT(t)t(1ν222)<0,a.s. (4.17)

    In other terms, T(t) approaches zero exponentially almost surely, such that the tumor will be completely eradicated from the community with unit probability.

    Proof. In order to prove the theorem, we will utilize the method of direct integration on the stochastic tumor-immune model system (2.2). Firstly, we will utilize Itˆo's formula to derive the second equation of model (2.2). This yields the following result:

    d(lnT(t))=[r(1βT)ξEν222]dt+ν2dW2[rν222]dt+ν2dW2. (4.18)

    On the integration of relation (4.18) over the interval 0 to t and subsequent division by t, the resulting expression is obtained as follows:

    lnT(t)t(rν222)+ν2W2(t)t+lnT(0)t. (4.19)

    We can conclude that based on the strong law of large numbers of Brownian motion [21],

    limtν2W2(t)t+lnT(0)t=0,a.s.

    By computing the limit superior of each side of the equation and since r<ν222

    limtsuplnT(t)t(rν222)<0,a.s. (4.20)

    It means that whenever T(t) goes to zero exponentially with probability one, then

    limtT(t)=0a.s.

    The proof is completed here.

    In this section, we explore the dynamics outlined in Eq (2.2) in the context of two distinct agents: an immune boost agent labeled ϑ1 and a chemotherapeutic agent labeled ϑ2. Like Paclitaxel, the first agent is commonly regarded as cytotoxic or apoptotic. In contrast, the second agent is viewed as a form of rudimentary immunotherapy, which involves the application of an interleukin-derived drug to stimulate the immune system. We aim to gain a deeper understanding of how different agents can be used to control and manage tumor growth within the immune system, to identify the most effective combination therapy strategies, and to gain insights into the efficacy of biotherapy and immune boosts. The tumor-immune model (2.2) can be described as follows:

    dE(t)=[αE(tτ1)T(tτ1)(η1ϑ2)E(t)+μ1E(t)I(t)+e1]dt+ν1E(t)dW1,dT(t)=[rT(1βT)(ξ+ϑ1)E(t)T(t)ϑ2T(t)]dt+ν2T(t)dW2,dI(t)=[(η2+ϑ1)E(tτ2)T(tτ2)μ2I(t)+e2]dt+ν3I(t)dW3. (5.1)

    We assume that the initial conditions for system (5.1) satisfy

    E(θ)=σ1(θ),T(θ)=σ2(θ),I(θ)=σ3(θ), θ[τ,0], τ=max{τ1,τ2}. (5.2)

    To facilitate clarity in our explanation, we shall establish a vector

    x(t)=[E(t),T(t),I(t)]andϑ(t)=[ϑ1,ϑ2]. (5.3)

    such that

    dx(t)=f(x(t),x(tτ),ϑ(t))dt+g(x(t))dw(t),fortτ,τ>0, (5.4)

    with the initial conditions

    x(0)=[E0,T0,I0]=x0C([τ,0];R3+). (5.5)

    where, f(,,) and g(,,) are two vectors, each of which has components such that

    f1(x(t),x(tτ),ϑ(t))=αE(tτ1)T(tτ1)(η1ϑ2)E+μ1EI+e1,f2(x(t),x(tτ),ϑ(t))=rT(1βT)(ξ+ϑ1)ETϑ2T,f3(x(t),x(tτ),ϑ(t))=(η2+ϑ1)E(tτ2)T(tτ2)μ2I+e2,

    and g1(x(t))=ν1E(t), g2(x(t))=ν2T(t), g3(x(t))=ν3I(t).

    Our goal is to minimize the number of tumor cells. As a result, the quadratic cost function is proposed and defined as follows:

    G(ϑ(t))=12E{tf0(K1E(t)+K2T(t)+K3I(t)+S12ϑ21(t)+S22ϑ22(t))dt+b12E2(t)+b22T2(t)+b32I2(t)}, (5.6)

    where Ki, Si and bi, for i=1,2,3 are positive constants. The aim of this section is to find an optimal control ϑ(t)=[ϑ1(t),ϑ2(t)] that possesses the following property:

    J(ϑ)J(ϑ),ϑU (5.7)

    where, the control set is denoted by the set U and is defined as follows:

    U={ϑi(t)|ϑi(t)[0,ϑmaxi],t(0,tf],ϑiL2[0,tf],i=1,2}, (5.8)

    where, ϑmaxiR+ are constants. Before applying the stochastic maximal criterion, it is necessary to determine the Hamiltonian Hm(p,q,ϑ,x) such that

    Hm(p,q,ϑ,x)=g(x),ql(x,ϑ)+f(x,ϑ),p, (5.9)

    where .,. represents the inner product in the Euclidean space; the vector fields f(x,ϑ),p,g(x),q and l(x,ϑ) are continuous functions on the inner product and continuously differentiable with repect to the state variable and the time variable; p=[p1,p2,p3] and q=[q1,q2,q3] refers to the two different sets of adjoint variables that are independent of one another. By employing a similar maximum criterion approach, we find:

    dx(t)=g(x(t))dW(t)+H(x,ϑ,p,q)pdt, (5.10)
    dp(t)=q(t)dW(t)xH(x,ϑ,p,q)dt, (5.11)
    Hm(x,ϑ,p,q)=minϑUHm(x,ϑ,p,q), (5.12)

    where, the state x(t) denotes the optimal trajectory followed by x(t). The initial condition at time t=0 for the state variables and the final condition (t=tf) for the adjoint variables p(t) of Eqs (5.10) and (5.11) are as follows:

    x(0)=x0andp(tf)=xh(x(tf)),

    respectively. The optimal value of the control variable ϑ can be expressed as a function of the adjoint variables p,q, and state x, as demonstrated by Eq (5.12):

    ϑ(t)=ϕ(x,q,p), (5.13)

    where the value of ϕ is determined by Eq (5.12). Thus, the Hamiltonian is represented by:

    H=K1E+K2T+K3I+S12ϑ21+S22ϑ22+b12E2+b22T2+b32I2+p1(αE(tτ1)T(tτ1)(η1ϑ2)E+μ1EI+e1)+p2(rT(1βT)(ξ+ϑ1)ETϑ2T)+p3((η2+ϑ1)E(tτ2)T(tτ2)μ2I+e2)+ν1Eq1+ν2Tq2+ν3Iq3. (5.14)

    Thus, we have

    ˙p1(t)=HEψ1[0,tfτ1]HE(tτ1)(t+τ1)ψ2[0,tfτ2]HE(tτ2)(t+τ2),˙p2(t)=HTψ1[0,tfτ1]HT(tτ1)(t+τ1)ψ2[0,tfτ2]HT(tτ2)(t+τ2),˙p3(t)=HI, (5.15)

    where

    ψ1[0,tfτ1]={1if t[0,tfτ1]0otherwise,ψ2[0,tfτ2]={1if t[0,tfτ2]0otherwise. (5.16)

    The Pontryagin maximum principle yields the adjoint system:

    dp1dt={K1+b1E+αp1T(tτ1)p1(η1ϑ2)+p1μ1Ip2(ξ+u1)T+p3(η2+ϑ1)T(tτ2)}+ν1q1,dp2dt={K2+b2T+αp1E(tτ1)+p2(r2rβT)p2(ξ+ϑ1)Ep2ϑ2+p3(η2+ϑ1)E(tτ2)}+ν2q2,dp3dt={K3+b3I+p1μ1Ep3μ2}+ν3q3, (5.17)

    where, p1(tf)=b1E(t),p2(tf)=b2T(t),p3(tf)=b3I(t). In addition to the condition of pi(tf)=0 at the final time tf, it holds true for values of i=1,2,3. Similarly, the supplementary initial and final conditions are: E(0)=ˆE,T(0)=ˆT,I(0)=ˆI, p(tf)=h(x(tf))x such that

    h(E,T,I)=b12E2+b22T2+b32I2. (5.18)

    Thus, the Hamiltonian function calculates the partial derivatives of ϑ1 and ϑ2, respectively; we get the following characterization of the OC

    ϑ1=max{min{1,(p2p3)ETS1},0}, (5.19)
    ϑ2=max{min{1,p2Ep1TS2},0}. (5.20)

    Remark 5.1. In ODEs/DDEs, time is continuous and the control inputs are typically continuous functions of time. The presence of randomness in stochastic differential equations (SDEs) makes optimal control more challenging, as the control input is required to account for the inherent uncertainty. In SDEs, optimal control policies may be stochastic, i.e., probability distributions over control inputs, rather than deterministic functions of time.

    The following section pertains to the numerical simulation of outcomes derived from a stochastic optimal control problem.

    In this section, numerical simulations are conducted to validate the main results derived previously. The results are validated through numerical simulations utilizing Milstein's higher order method [39], which exhibits a strong order of convergence one. This method is employed to numerically solve the stochastic model (2.2). Subsequently, the discretization system that corresponds to the aforementioned method is obtained.

    dEn+1=En+h[αEnm1Tnm1η1En+μ1EnIn+e1]+ν1Enχ1,nh+ν212En[χ21,n1]h,dTn+1=Tn+h[rTn(1βTn)ξEnTn]+ν2Tnχ2,nh+ν222Tn[χ22,n1]h,dIn+1=In+h[η2Enm2Tnm2μ2In+e2]+ν3Inχ3,nh+ν232In[χ23,n1]h, (6.1)

    where the mutually independent N(0,1) random variables are denoted by χ1,n, χ2,n and χ3,n, the integers m1 and m2 make the time-delays τ1 and τ2 to be represented with the step size h, with τ1=m1h and τ2=m2h, respectively.

    Consider the initial conditions of the system (2.2) as E(0)=3.8,T(0)=1.0,I(0)=0.5, and the parameters are presented in Table 1. For the following numerical results, the time unit is days.

    Based on the parameter values taken from Table 1, Figure 1 displays the phase and time figures of the system (2.1). Theorem 7.1 states that system (2.1) has a unique equilibrium point that is locally asymptotically stable. Figure 1(a) illustrates that all trajectories converge towards the equilibrium point S=(E,T,I). Figure 1(b) illustrates that the sizes of three distinct cell types remain constant following a period of fluctuation.

    Figure 1.  Shows, a positive unique equilibrium state, S=(0.9002,0.0350,6.9011) is locally asymptotically stable, with initial values (3.2, 1.2, 0.7), (3.4, 1.4, 0.9), (3.6, 1.6, 0.9), (3.8, 1.8, 1.1), (4.0, 2.0, 1.3), (4.2, 2.2, 1.5).

    Figure 2 shows the simulations results of (E(t),T(t),I(t)) for stochastic model (2.2) with its corresponding deterministic model (2.1). According to Theorem 4.1 and Figure 3, the tumor size T(t) in the system (2.2) is persistent. It demonstrates that the model has a unique stationary distribution and that the disease is chronic, resulting in relatively low white noise intensities where ˆT0>1, where the white noise disturbance level of system states is νi=0.001,i=1,2,3. The solution of system (2.2) approaches equilibrium S and is stochastic asymptotically stable, as can be seen in Figure 3.

    Figure 2.  Trajectories of (E(t),T(t),I(t)) of stochastic model with its deterministic version (2.1).
    Figure 3.  The model (2.2)'s numerical simulations demonstrate that the system exhibits a distinct ergodic stationary distribution in cases where the tumor is persistent and ˆT0 exceeds 1, with νi values of 0.001 for i = 1, 2 and 3.

    The efficacy of cancer treatments may vary from patient to patient or even within a single patient over time. When this variability is taken into account, white noise can be introduced into the model. A stochastic model refers to white noise as random fluctuations or variability. Different factors can cause it, and increasing the white noise levels can have interesting effects on the system's dynamics, such as accelerating tumor cell death. Increasing white noise in a model can be caused by external factors that introduce randomness or fluctuations into the environment.

    Tumor growth can be influenced by fluctuations in nutrient availability, temperature, or oxygen levels, for example in Figure 4, we simulate the stochastic model (2.2) with different white noise levels. The results in Theorem 4.2 are validated when the white noise level is increased (see Figure 4(b)). As shown in Figure 4(a), the effector cells increase with the level of white noise. As shown in Figure 5(a), the histogram represents the population of tumor cells under the influence of white noise disturbance intensities ν2=0.02. On the other hand, Figures 5(b), (c) illustrate that tumor cell population decreases as noise disturbance intensities increase, as ν2=0.1 and ν3=0.15, respectively.

    Figure 4.  Shows the effects of different white noise disturbance levels on the stochastic model (2.2).
    Figure 5.  The histogram diagram of tumor size I(t). The white noise disturbances of (a) is ν2=0.07, (b) is ν2=0.1 and (c) is ν2=0.15.

    The graphical findings in Figure 6 demonstrate the biological importance of delay parameters. It is possible to compare tumor immune interactions with different delays based on these results. There is a significant correlation between delays and stochastic tumor immune system as shown in Figure 6. Compared to the model with time delay τ1=0.4,τ=0.38 with a small value of delay (τ1=0.05, τ2=0.1), the plot of model (2.2) with small value of delay (τ1=0.05,τ2=0.1) displays a significant deviation. Thus, delays significantly impact the analysis of stochastic tumor immune interactions' dynamic characteristics.

    Figure 6.  Displays the effects of different values of time delays on the stochastic model (2.2).

    OC is an essential mechanism in mathematical modeling that targets to figure out the best appropriate control procedure for a given system, based on particular constraints. Therefore, OC can be used to outline medication protocols that minimize the growth of tumors and maximize the stimulating of the immune system, while considering the random fluctuations that arise in biological processes. To numerically solve the stochastic OC-P:

    ● Discretize the stochastic model with time-delays using a numerical scheme such as Milstein's higher order method.

    ● Carry out the forward simulations which includes the discretized stochastic model, with a particular set of control variables.

    ● Include computing the objective function at every step of the forward approximation.

    ● Execute the backward approximation by including the adjoint system in order to find the gradient of the objective function with regard to the input variables.

    ● Applying an optimization method, such as the stochastic approximation scheme, to minimize the objective function based on the system solutions and control constraints, and then modify the control inputs accordingly.

    The numerical simulations of the optimal control problem (5.1), along with the adjoint equation (5.17) and the characterization of the optimal control (5.19), (5.20), are presented graphically. According to Table 1, the graphic outcomes of an optimal control problem were obtained by comparing them with the results of a scenario without a control system. Figure 7 shows that the controls employed are highly effective in eliminating the disease. Figure 7(a)(f) illustrate the time evolution of effector, tumor, and IL-2 cells without and with controls ϑ1 and ϑ2. Figure 7(a)(f) demonstrate the efficacy of the implemented controls in reducing the number of tumor cells and maximizing the number of effector cells. In the absence of chemotherapy and an immune boost, the tumor cell population exhibits an increasing trend over time, whereas the presence of treatment helps the immune system's ability to regulate the proliferation of tumor cells. Evidently, as the immune boost and chemotherapeutic agent are enhanced, there is a corresponding rise in effector cells. In contrast, tumor cells lead to their decline and eventual extinction. Meanwhile, IL-2 cells tend to return to baseline levels.

    Figure 7.  Numerical simulations of the OC-P for ODE and SDE (2.1), (2.2), before and after OC which reduces the growth of the tumor cells.

    Tumor modeling requires understanding the effects of environmental changes such as changes in nutrient availability, temperature, and oxygen levels on tumor-immune interactions. By incorporating discrete time-delay parameters and multiplicative white noise terms, the current work investigates the dynamical behavior of tumor-immune interactions with external treatments. The influence of environmental noises on the persistence and possible extinction of tumor cells has been studied under circumstances where the intensity of randomly varying driving forces is extremely diverse. A combination of white noises and time delays greatly affected the dynamics of the tumor-immune interaction model. It has been seen that, the presence of stochastic perturbations with a relatively small scale of white noise, tumor cells oscillate within a wide range of values, whereas large noises can lead to the eradication of the tumor cells.

    To minimize tumor cells and maximize effector cells and IL-2 concentration, control variables are included in the stochastic model. An optimal control problem has been investigated to manage tumor growth and identify the most effective combination therapy strategies, and gain insights into the efficacy of biotherapy and immune boosts. Numerical results demonstrate optimality in the control variables and the effectiveness of introducing additional controls to reduce the growth of tumor cells. As a result of the presence of white noise in the optimality problem, tumor growth can be reduced.

    A color-coded noise, such as telegraph noise, can be described as a random switch between two or more regimes of environmental influence on tumor immune response [40]. Future studies will examine how regime switching affects tumor cells, effector cells, and IL-2 cells. A sensitivity analysis can also be used to analyze tumor-free steady-state stability under small parameter variations.

    The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.

    This work has been generously supported by UAE-ZU project #12S107.

    The authors declare there is no conflict of interest.

    The equilibrium states of the system (2.1) are:

    (a) The tumor-free steady state S, refers to a condition in which the population of tumor cells is absent, while the normal cells stay healthy:

    S=(E,T,I)=(e1μ2η1μ2μ1e2,0,e2μ2).

    (b) The tumor-infection steady state S:

    S=(E,T,I)=(r(1βT)ξ,T,e2+η2r(1βT)Tξμ2)

    where the T represents the positive solutions of the below equation:

    e1+αr(1βT)Tξη1r(1βT)ξ+μ1r(1βT)ξ[e2+η2r(1βT)Tξμ2]=0.

    In order to analyze the stability of the equilibrium states, it is common practice to linearize the system around the equilibrium states and determine the Jacobian matrices. Currently, our focus is only on examining the stability of the equilibrium point in the presence of a tumor.

    The Jacobian matrix ΔS(E,T,I) is obtained in order to determine the eigenvalues of the system (2.1).

    ΔS=(αTeλτ1η1+μ1IαEeλτ1μ1EξTr(12βT)ξE0η2Teλτ2η2Eeλτ2μ2). (7.1)

    Remark 7.1. The stability analysis of the equilibrium states is examined by the eigenvalues of ΔS. If the real parts of all the eigenvalues of the matrix ΔS are negative, then the equilibrium state S exhibits local asymptotic stability. Conversely, if at least one of the eigenvalues exhibit a positive real part, the equilibrium state is considered to be unstable.

    Following some calculations, the characteristic equation of (7.1) is obtained:

    Det(ΔS)=B1(λ)+eλτ1B2(λ)+eλτ2B3(λ)=0 (7.2)

    where,

    B1(λ)=λ3+w1λ2+w2λ+w3,B2(λ)=λ2w4+λw5+w6,B3(λ)=w7,w1=η1+μ2μ1I+2βrTr+ξE,w2=μ1I(μ2+2βrTr+ξE)+η1(μ2+2βrTr+ξE)+μ2(2βrTr+ξE),w3=μ2(η1μ1I)(r(2βT1)+ξE),w4=αT,w5=2αβr(T)2+αrTαμ2T,w6=2αβμ2r(T)2+αμ2rT,w7=2βη2μ1r(T)2E+η2μ1rTEη2λμ1TE.

    Case (ⅰ): τ1=0,τ2=0, we have

    Det(ΔS)=λ3+(w1+w4)λ2+(w2+w5)λ+(w3+w6+w7)=0. (7.3)

    Based on the Routh-Hurwitz criterion, all the eigenvalues of the characteristic equation Det(ΔS)=0 have negative real parts if the following conditions hold:

    w1+w4>0,w3+w6+w7>0,(w1+w4)(w2+w5)>(w3+w6+w7). (7.4)

    Therefore, it can be concluded that the asymptotic stability of the system (2.1) is achieved in the equilibrium state S, when τ1=τ2=0.

    Case (ⅱ): τ1=τ2>0, we have

    Det(ΔS)=λ3+w1λ2+w2λ+w3+eλτ(λ2w4+λw5+w6+w7)=0. (7.5)

    Assume that the pure imaginary root λ=iθ,θ>0. Then applying the λ=iθ into (7.5) and separating its real and imaginary parts, we have

    {w3w1θ2+(w6+w7w4θ2)cos(θτ)+w5θsin(θτ)=0,w2θθ3+w5θcos(θτ)(w6+w7w4θ2)sin(θτ)=0. (7.6)

    Through mathematical calculations, the following equations are derived:

    cos(θτ)=θ4(w5w1w4)+θ2(w3w4w2w5+w1w6+w1w7)w3w6w3w7θ4w24+θ2(w252w4w62w4w7)+w26+w27+2w6w7, (7.7)
    sin(θτ)=θ5w4+θ3(w2w4+w1w5w6w7)+θ(w3w5+w2w6+w2w7)θ4w24+θ2(w252w4w62w4w7)+w26+w27+2w6w7. (7.8)

    Using the trigonometric formula cos2(θτ)+sin2(θτ)=1, we have

    θ10+θ8h1+θ6h2+θ4h3+θ2h4+h5=0 (7.9)

    where,

    h1=(w44+w21w242w2w242w6w42w7w4+w25)/w24,h2=(4w6w34+4w7w34+w22w242w25w242w1w3w242w21w6w4+4w2w6w42w21w7w4+4w2w7w4+w21w252w2w25+w26+w27+2w6w7)/w24,h3=(w45+w22w252w1w3w25+4w4w6w25+4w4w7w25+w23w24+w21w266w24w262w2w26+w21w276w24w272w2w272w22w4w6+4w1w3w4w62w22w4w7+4w1w3w4w7+2w21w6w712w24w6w74w2w6w7)/w24,h4=4(w4w36+w22w262w25w262w1w3w26+12w4w7w26+12w4w27w62w23w4w6+2w22w7w64w25w7w64w1w3w7w6+4w4w37+w23w25+w22w272w25w272w1w3w272w23w4w7)/w24,h5=(w46w474w6w37+w23w26+w23w276w26w274w36w7+2w23w6w7)/w24.

    Therefor, the equilibrium states of system (2.1) are asymptotically stable if and only if (7.9) has no positive real roots. Assume that ω=θ2, we have

    ω5+ω4h1+ω3h2+ω2h3+ωh4+h5=0. (7.10)

    If hp<0,p=1,,5, then (7.10) has at least one positive root and if hp>0,p=1,,5, then (7.10) has all the roots must be negative.

    Case (ⅲ): τ1=0,τ2>0. Then, (7.2) can be written as

    B1(s)+B2(s)+eλτ2B3(s)=0. (7.11)

    Let suppose λ=iω is a pure imaginary root of (7.11), then

    {cosωτ2w7=(w1+w4)ω2(w3+w6),sinωτ2w7=(w1+w5)ω3. (7.12)

    It can be deduced from Eq (7.12) that,

    ω6+z1ω4+z2ω2+z3=0 (7.13)

    where,

    z1=a212a3,z2=a232a1a2,z3=a22w27.

    Denote β=ω2, (7.13) becomes

    β3+z1β2+z2β+z3=0. (7.14)

    Then let

    G(ω)=β3+z1β2+z2β+z3. (7.15)

    If the condition (C1)z3<0 is satisfied, given that dF(ω)dω>0 ω>0, it can be concluded that Eq (7.13) possesses at least one positive real root. Therefore, Eq (7.11) has at least one pair of purely imaginary roots. According to the findings of Sun et al. [42], the subsequent outcome is as follows:

    Lemma 7.1. The following holds true for (7.11):

    (a) If condition (7.4) is satisfied and the values of z1,z2,z3>0, then equation (7.11) does not has any roots with zero real parts for values of τ20.

    (b) If z3<0 and z1,z2>0 holds, then (7.11) has a pair of purely imaginary roots ±iω0 when τ2=τ2j and for any ω0 (unique positive zero of the function G(ω))

    τ2j=1ω0(arccos[a1ω2a2w7]+2jπ). (7.16)

    The proof of Lemma 7.1 has resemblance to the proof of Lemma 3.1 as presented in the work of Li et al. [43]. Here, it is left out.

    (C2) ˜A1˜A2+˜B1˜B2>0

    where,

    ˜A1=3ω20+w2+2ω0w4sinω0τ1+w5cosω0τ1,˜A2=ω20w5,˜B1=2ω0w1+2ω0w4cosω0τ1sinω0τ1w5,˜B2=ω30w4+w6ω0.

    Case (ⅳ): When τ1>0,τ2=0, the Eq (7.2) becomes

    B1(s)+B3(s)+eλτ1B2(s)=0. (7.17)

    For pure imaginary root λ=iω, Eq (7.17) can be written as

    {cosωτ1ωw5sinωτ1w6+sinωτ1ω2w4ω3+ωw2=0,cosωτ1(w6+ω2w4)+sinωτ1ωw5w1ω2+w3+w7=0. (7.18)

    It can be deduced from Eq (7.18) that,

    ω10+r1ω8+r2ω6+r3ω4+r4ω2+r5=0. (7.19)

    where,

    r1=(w44+w21w242w2w244w1w5w4+2w6w4+w25)/w24,r2=(w25w212w4w6w212w3w24w1+4w2w4w5w12w24w7w1+w22w24+2w24w252w2w25+w26+4w3w4w54w2w4w6+4w4w5w7)/w24,r3=(w45+w22w252w1w3w252w1w7w254w2w3w4w54w2w4w7w5+w23w24+w21w26+2w24w262w2w26+w24w27+2w22w4w6+4w1w3w4w6+2w3w24w7+4w1w4w6w7)/w24,r4=(w25w232w4w6w232w1w26w3+2w25w7w34w4w6w7w3+w22w262w25w26+w25w272w4w6w272w1w26w7)/w24,r5=(w2w46+w23w26+w26w27+2w3w26w7)/w24.

    Take α=ω2, then (7.19) becomes

    ω5+r1ω4+r2ω3+r3ω2+r4ω+r5=0 (7.20)

    and let

    F(ω)=α5+r1α4+r2α3+r3α2+r4α+r5. (7.21)

    If the condition (C3)r5<0 is satisfied, given that dF(ω)dω>0 ω>0, it can be concluded that Eq (7.20) possesses at least one positive real root. Therefore, Eq (7.17) has at least one pair of purely imaginary roots. According to the findings of Sun et al. [42], the subsequent outcome is as follows:

    Lemma 7.2. The following holds true for (7.17):

    (a) If condition (7.4) is satisfied and the values of ri>0,i=1,2,3,4,5, then Eq (7.17) does not have any roots with zero real parts for values of τ10.

    (b) If r5<0 and ri>0,i=1,2,3,4 holds, then (7.17) has a pair of purely imaginary roots ±iω0 when τ1=τ1j and for any ω0 (unique positive zero of the function F(ω))

    τ1j=1ω0(arccos[(w2w4w6)(w1w2w3w7)w(w3ww2)w5(w2w4w6)(w4w2+w6)w2w25]+2jπ). (7.22)

    The proof of Lemma 7.2 has resemblance to the proof of Lemma 3.1 as presented in the work of Li et al. [43]. Here, it is left out.

    (C4) ˆA1ˆA2+ˆB1ˆB2>0

    where,

    ˆA1=3ω20+(w2+w5),ˆA2=ω0sinω0τ0w7,ˆB1=2ω0(w1+w4),ˆB2=ω0cosω0τ0w7.

    Case (ⅴ): τ1>0,τ2>0. Let the delay τ2[0,τ20) in (7.2) and choose τ1 as a bifurcating parameter. Thus let pure imaginary root (7.2) as λ=iω=ω(cosπ2+sinπ2). Then (7.2) becomes

    {cosωτ1(w6w4ω2)+sinωτ1ωw5=ω2w1+w3+cosωτ2w7,cosωτ1ωw5+sinωτ1(w4ω2w6)=ωw2+ω3+sinωτ2w7. (7.23)

    By (7.23), we have

    cosωτ2=(cosωτ1(w6w4ω2)+sinωτ1ωw5ω2w1w3)/w7, (7.24)
    sinωτ2=(cosωτ1ωw5+sinωτ1(w4ω2w6)+ωw2ω3)/w7. (7.25)

    In view of cos2ωτ2+sin2ωτ2=1, we have

    ω6+q1ω5+q2ω4+q3ω3+q4ω2+q5ω+q6=0, (7.26)

    where

    q1=2w4sin(ωτ1),q2=(w24sin2(ωτ1)+w24cos2(ωτ1)+2w4w1cos(ωτ1)2w5cos(ωτ1)+w212w2),q3=(2w2w4sin(ωτ1)2w1w5sin(ωτ1)+2w6sin(ωτ1)),q4=(w25sin2(ωτ1)2w4w6sin2(ωτ1)+w25cos2(ωτ1)2w4w6cos2(ωτ1)2w3w4cos(ωτ1)+2w2w5cos(ωτ1)2w1w6cos(ωτ1)+w222w1w3),q5=(2w3w5sin(wy1)2w2w6sin(wy1)),q6=w26sin2(ωτ1)+w26cos2(wy1)+2w3w6cos(wy1)+w23w27.

    Then let the polynomial function

    H(ω)=ω6+q1ω5+q2ω4+q3ω3+q4ω2+q5ω+q6. (7.27)

    When the condition (C5) q6<0 holds, given that dH(ω)dω>0 ω>0, it can be concluded that Eq (7.26) possesses at least one positive real root. Therefore, Eq (7.2) has at least one pair of purely imaginary roots. According to the findings of Sun et al. [42], the subsequent outcome is as follows:

    Lemma 7.3. The following holds for (7.2):

    (ⅰ) If qi>0,i=1,2,,6, then Eq (7.2) has no roots with zero real parts.

    (ⅱ) If the value of q6 is less than zero, it can be concluded that Eq (7.2) possesses at least one pair of purely imaginary roots.

    (ⅲ) When given that q6>0 and there exists a constant δ>0 such that H(δ)<0, it can be concluded that Eq (7.2) possesses at least two pairs of roots that are totally imaginary.

    The proof of Lemma 7.3 has resemblance to the proof of Lemma 3.1 as presented in the work of Li et al. [44]. Here, it is left out. Now, it is assumed that Eq (7.26) possesses a positive real root ω. By (7.24), we have

    τ2j=1ω[arccos((cosωτ1(w6w4ω2)+sinωτ1ωw5ω2w1w3)/w7)+2jπ], (7.28)

    where, j=0,1,2,3,. Then ±iω is a pair of roots of (7.2) when τ2=τ2j.

    (C6) ˉA1ˉA2+ˉB1ˉB2>0

    where,

    ˉA1=3ω20+w2+w5cosω0τ1+2ω0w4sinω0τ1+τ1ω20w4cosω0τ1τ1w6cosω0τ1τ1sinω0τ1ω0w5,ˉB1=2ω0w1+2ω0w4cosω0τ1w5sinω0τ1τ1ω0w5cosω0τ1τ1ω20w4sinω0τ1+τ1w6sinω0τ1,ˉA2=ω0w7sinω0τ2,ˉB2=ω0w7cosω0τ2.

    Now, we reach the following results:

    Theorem 7.1. If system (2.1) has an tumor-infection equilibrium state S, then the following statements are hold:

    (ⅰ) For τ1=τ2=0, the equilibrium state S of system (2.1) is asymptotically stable if the conditions in (7.4) are true.

    (ⅱ) For τ1=τ20, the tumor-infection equilibrium state S is asymptotically stable if all the roots of Det(ΔS)=0 are negative real parts.

    (ⅲ) If τ1=0 and τ2>0 and (7.4), (C1),(C2) hold, then the tumor-infection equilibrium state S is asymptotically stable when τ2[0,τ20).

    (ⅳ) If τ2=0 and τ1>0 and (7.4), (C3),(C4) hold, then the tumor-infection equilibrium state S is asymptotically stable when τ1[0,τ10).

    (ⅴ) If τ1[0,τ10), and (7.4), (C5) and (C6) are satisfied, then the tumor-infection equilibrium state S is asymptotically stable when τ2[0,τ20).

    Remark 7.2. The stability of the model (2.1) is completely determined by the zeros of its characteristic polynomial (7.2). To find the number of zeros of (7.2) as time delays τ1 and τ2 vary, we write Det(ΔS) in (7.2) as ΔS(λ,τ1,τ2). In this case the solution of the system is stable, the coefficient polynomials Λ1(λ)=λ3+w1λ2+w2λ+w3,Λ2(λ)=λ2w4+λ+w5+w6,Λ3(λ)=w7 satisfies the following conditions[41]: (1) deg(Λ1(λ))max{deg(Λ2(λ)),deg(Λ3(λ))}, (2) Λ1(0)+Λ2(0)+Λ3(0)0, (3) The polynomials Λ1(λ),Λ2(λ) and Λ3(λ) do not have any common zeros, (4) limλ(|Λ2(λ)/Λ1(λ)|+|Λ3(λ)/Λ1(s)|)<1. In the absence of condition (1), the polynomial Det(ΔS) is not stable for any positive delays. If condition (2) is not satisfied, then the polynomial Det(ΔS) has 0 as a root for any τ1 and τ2, thus, it can never be stable. Condition (3) is natural. (4) is automatically satisfied since its left-hand side is zero. In condition (5) the number of zeros of Det(ΔS) can change only if a zero appears on the imaginary axis.



    [1] World Health Organization, Cancer, 2022. Available from: https://www.who.int/news-room/fact-sheets/detail/cancer
    [2] A. Desai, T. Mohammed, N. Duma, M. Garassino, L. Hicks, N. Kuderer, et al., COVID-19 and cancer: A review of the registry-based pandemic response, JAMA Oncol., 7 (2021), 1882–1890. https://doi.org/10.1001/jamaoncol.2021.4083 doi: 10.1001/jamaoncol.2021.4083
    [3] K. Dehingia, H. Sarmah, Y. Alharbi, K. Hosseini, Mathematical analysis of a cancer model with time-delay in tumor-immune interaction and stimulation processes, Adv. Differ. Equation, 2021 (2021), 1–27. https://doi.org/10.1186/s13662-021-03621-4 doi: 10.1186/s13662-021-03621-4
    [4] F. A. Rihan, K. Udhayakumar, Fractional order delay differential model of a tumor-immune system with vaccine efficacy: Stability, bifurcation and control, Chaos Solitons Fractals, 173 (2023) 113670. https://doi.org/10.1016/j.chaos.2023.113670 doi: 10.1016/j.chaos.2023.113670
    [5] F. A. Rihan, G. Velmurugan, Dynamics of fractional-order delay differential model for tumor-immune system, Chaos Solitons Fractals, 132 (2020), 109592. https://doi.org/10.1016/j.chaos.2019.109592 doi: 10.1016/j.chaos.2019.109592
    [6] V. Bitsouni, V. Tsilidis, Mathematical modeling of tumor-immune system interactions: The effect of rituximab on breast cancer immune response, J. Theor. Biol., 539 (2022), 111001. https://doi.org/10.1016/j.jtbi.2021.111001 doi: 10.1016/j.jtbi.2021.111001
    [7] M. Itik, S. Banks, Chaos in a three-dimensional cancer model, Int. J. Bifurcat. Chaos, 20 (2010), 71–79. https://doi.org/10.1142/S0218127410025417 doi: 10.1142/S0218127410025417
    [8] R. Yafia, A study of differential equation modeling malignant tumor cells in competition with immune system, Int. J. Biomath., 4 (2011), 185–206. https://doi.org/10.1142/S1793524511001404 doi: 10.1142/S1793524511001404
    [9] Y. Radouane, Hopf bifurcation in a delayed model for tumor-immune system competition with negative immune response, Discrete Dyn. Nat. Soc., 2006 (2006), 095296. https://doi.org/10.1155/DDNS/2006/95296 doi: 10.1155/DDNS/2006/95296
    [10] F. Najm, R. Yafia, M. A. Aziz-Alaoui, Hopf bifurcation in oncolytic therapeutic modeling: Viruses as anti-tumor means with viral lytic cycle, Int. J. Bifurcat. Chaos, 32 (2022), 2250171. https://doi.org/10.1142/S0218127422501711 doi: 10.1142/S0218127422501711
    [11] R. Brady, H. Enderling, Mathematical models of cancer: When to predict novel therapies, and when not to, Bull. Math. Biol., 81 (2019), 3722–3731. https://doi.org/10.1007/s11538-019-00640-x doi: 10.1007/s11538-019-00640-x
    [12] T. Phan, S. Crook, A. Bryce, C. Maley, E. Kostelich, Y. Kuang, Mathematical modeling of prostate cancer and clinical application, Appl. Sci., 10 (2020), 2721. https://www.mdpi.com/2076-3417/10/8/2721
    [13] O. Nave, Adding features from the mathematical model of breast cancer to predict the tumour size, Int. J. Comput. Math. Comput. Syst. Theory, 5 (2020), 159–174. https://doi.org/10.1080/23799927.2020.1792552 doi: 10.1080/23799927.2020.1792552
    [14] D. Kirschner, J. Panetta, Modeling immunotherapy of the tumor-immune interaction, J. Math. Biol., 37 (1998), 235–252. https://doi.org/10.1007/s002850050127 doi: 10.1007/s002850050127
    [15] V. Kuznetsov, L. Makalkin, M. Taylor, A. Perelson, Nonlinear dynamics of immunogenic tumors: Parameter estimation and global bifurcation analysis, Bull. Math. Biol., 56 (1994), 295–321. https://doi.org/10.1016/S0092-8240(05)80260-5 doi: 10.1016/S0092-8240(05)80260-5
    [16] A. Omame, C. Nnanna, S. Inyama, Optimal control and cost-effectiveness analysis of an HPV-chlamydia trachomatis co-infection model, Acta Biotheor., 69 (2021), 185–223. 10.1007/s10441-020-09401-z doi: 10.1007/s10441-020-09401-z
    [17] U. Ijeoma, S. Inyama, A. Omame, Mathematical model and optimal control of new-castle disease (ND), Appl. Math. Comput., 9 (2020), 70–84. doi: 10.11648/j.acm.20200903.14 doi: 10.11648/j.acm.20200903.14
    [18] F. A. Rihan, S. Lakshmanan, H. Maurer, Optimal control of tumor-immune model with time-delay and immuno-chemotherapy, Appl. Math. Comput., 353 (2019), 147–165. https://doi.org/10.1016/j.amc.2019.02.002 doi: 10.1016/j.amc.2019.02.002
    [19] F. A. Rihan, H. J. Alsakaji, S. Kundu, O. Mohamed, Dynamics of a time-delay differential model for tumor-immune interactions with random noise, Alex. Eng. J., 61 (2022), 11913–11923. https://doi.org/10.1016/j.aej.2022.05.027 doi: 10.1016/j.aej.2022.05.027
    [20] M. Yu, Y. Dong, Y. Takeuchi, Dual role of delay effects in a tumour–immune system, J. Biol. Dyn., 11 (2017), 334–347. https://doi.org/10.1080/17513758.2016.1231347 doi: 10.1080/17513758.2016.1231347
    [21] X. Mao, Stochastic Differential Equations and Applications, Elsevier, 2007. https://doi.org/10.1016/B978-1-904275-34-3.50014-1
    [22] M. Baar, L. Coquille, H. Mayer, M. Hölzel, M. Rogava, T. Tüting, et al., A stochastic model for immunotherapy of cancer, Sci. Rep., 6 (2016), 1–10. https://doi.org/10.1038/srep24169 doi: 10.1038/srep24169
    [23] L. Han, C. He, Y. Kuang, Dynamics of a model of tumor-immune interaction with time delay and noise, DCDS-S, 13(2020). http://dx.doi.org/10.3934/dcdss.2020140 doi: 10.3934/dcdss.2020140
    [24] H. J. Alsakaji, F. A. Rihan, A. Hashish, Dynamics of a stochastic epidemic model with vaccination and multiple time-delays for COVID-19 in the UAE, Complexity, 2022 (2022), 1–15. https://doi.org/10.1155/2022/4247800 doi: 10.1155/2022/4247800
    [25] C. Odoux, H. Fohrer, T. Hoppo, L. Guzik, D. Stolz, D. Lewis, et al., A stochastic model for cancer stem cell origin in metastatic colon cancer, Cancer Res., 68 (2008), 6932–6941. https://doi.org/10.1158/0008-5472.CAN-07-5779 doi: 10.1158/0008-5472.CAN-07-5779
    [26] Y. Deng, M. Liu, Analysis of a stochastic tumor-immune model with regime switching and impulsive perturbations, Appl. Math. Model., 78 (2020), 482–504. https://doi.org/10.1016/j.apm.2019.10.010 doi: 10.1016/j.apm.2019.10.010
    [27] A. Raza, J. Awrejcewicz, M. Rafiq, N. Ahmed, M. Mohsin, Stochastic analysis of nonlinear cancer disease model through virotherapy and computational methods, Mathematics, 10 (2022), 368. https://doi.org/10.3390/math10030368 doi: 10.3390/math10030368
    [28] K. Dehingia, H. Sarmah, K. Hosseini, K. Sadri, S. Salahshour, C. Park, An optimal control problem of immuno-chemotherapy in presence of gene therapy, AIMS Math., 6 (2021), 11530–11549. https://doi.org/10.3934/math.2021669 doi: 10.3934/math.2021669
    [29] F. A. Rihan, Delay Differential Equations and Applications to Biology, Springer, 2021. https://doi.org/10.1007/978-981-16-0626-7
    [30] C. Orrieri, E. Rocca, L. Scarpa, Optimal control of stochastic phase-field models related to tumor growth, ESAIM Control Optim. Calc. Var., 26 (2020), 104. https://doi.org/10.1051/cocv/2020022 doi: 10.1051/cocv/2020022
    [31] M. Huang, S. Liu, X. Song, X. Zou, Control strategies for a tumor-immune system with impulsive drug delivery under a random environment, Acta Math. Sci., 42 (2022), 1141–1159. https://doi.org/10.1007/s10473-022-0319-1 doi: 10.1007/s10473-022-0319-1
    [32] L. J. Allen, An Introduction to Stochastic Processes with Applications to Biology, CRC press, 2010. https://doi.org/10.1201/b12537
    [33] F. Rihan, H. Alsakaji, Persistence and extinction for stochastic delay differential model of prey predator system with hunting cooperation in predators, Adv. Differ. Equations, 2020 (2020), 1–22 https://doi.org/10.1186/s13662-020-02579-z doi: 10.1186/s13662-020-02579-z
    [34] X. Mao, C. Yuan, Stochastic Differential Equations with Markovian Switching, World Scientific, 2006. https://doi.org/10.1142/p473
    [35] R. Hasminskii, Stochastic Stability of Differential Equations, Springer-Verlag Berlin Heidelberg, 2012. https://doi.org/10.1007/978-3-642-23280-0
    [36] S. Rajasekar, M. Pitchaimani, Qualitative analysis of stochastically perturbed SIRS epidemic model with two viruses, Chaos Solitons Fractals, 118 (2019), 207–221. https://doi.org/10.1016/j.chaos.2018.11.023 doi: 10.1016/j.chaos.2018.11.023
    [37] E. Beretta, V. Kolmanovskii, L. Shaikhet, Stability of epidemic model with time delays influenced by stochastic perturbations, Math. Comput. Simul., 45 (1998), 269–277. https://doi.org/10.1016/S0378-4754(97)00106-7 doi: 10.1016/S0378-4754(97)00106-7
    [38] M. Kinnally, Stationary Distributions for Stochastic Delay Differential Equations with Non-negativity Constraints, University of California, San Diego, 2009.
    [39] G. Milstein, Numerical Integration of Stochastic Differential Equations, Springer Science & Business Media, 1994. https://doi.org/10.1007/978-94-015-8455-5
    [40] Q. Luo, X. Mao, Stochastic population dynamics under regime switching, J. Math. Anal. Appl., 334 (2007), 69–84. https://doi.org/10.1016/j.jmaa.2006.12.032 doi: 10.1016/j.jmaa.2006.12.032
    [41] Q. An, E. Beretta, Y. Kuang, C. Wang, H. Wang, Geometric stability switch criteria in delay differential equations with two delays and delay dependent parameters, J. Differ. Equation, 266 (2019), 7073–7100. https://doi.org/10.1016/j.jde.2018.11.025 doi: 10.1016/j.jde.2018.11.025
    [42] Q. Sun, M. Xiao, M. B. Tao, Local bifurcation analysis of a fractional-order dynamic model of genetic regulatory networks with delays, Neural Process. Lett., 47 (2018), 1285–1296. https://doi.org/10.1007/s11063-017-9690-7 doi: 10.1007/s11063-017-9690-7
    [43] L. Li, Z. Wang, Y. Li, H. Shen, J. Lu, Hopf bifurcation analysis of a complex-valued neural network model with discrete and distributed delays, Appl. Math. Comput., 330 (2018), 152–169. https://doi.org/10.1016/j.amc.2018.02.029 doi: 10.1016/j.amc.2018.02.029
    [44] C. Xu, M. Liao, P. Li, Y. Guo, Q. Xiao, S. Yuan, Influence of multiple time delays on bifurcation of fractional-order neural networks, Appl. Math. Comput., 361 (2019), 565–582. https://doi.org/10.1016/j.amc.2019.05.057 doi: 10.1016/j.amc.2019.05.057
  • This article has been cited by:

    1. Girija Panneerselvam, Prakash Mani, Non-fragile event-triggered control for PMSM model with stochastic disturbances, 2024, 1951-6355, 10.1140/epjs/s11734-024-01199-y
    2. Anwarud Din, Yongjin Li, Optimizing HIV/AIDS dynamics: stochastic control strategies with education and treatment, 2024, 139, 2190-5444, 10.1140/epjp/s13360-024-05605-1
    3. Yan Fu, Tian Lu, Meng Zhou, Dongwei Liu, Qihang Gan, Guowei Wang, Effect of color cross-correlated noise on the growth characteristics of tumor cells under immune surveillance, 2023, 20, 1551-0018, 21626, 10.3934/mbe.2023957
    4. Marya Sadki, Karam Allali, Stochastic two-strain epidemic model with saturated incidence rates driven by Lévy noise, 2024, 375, 00255564, 109262, 10.1016/j.mbs.2024.109262
    5. Ali Sadiq Alabdrabalnabi, Ishtiaq Ali, Stability analysis and simulations of tumor growth model based on system of reaction-diffusion equation in two-dimensions, 2024, 9, 2473-6988, 11560, 10.3934/math.2024567
    6. P. Muthukumar, K. Anukiruthika, Optimal control system of multi-term fractional stochastic inclusion with Clarke’s subdifferential, 2024, 1951-6355, 10.1140/epjs/s11734-024-01200-8
    7. Mohammed Salman, Pradeep Kumar Das, Sanjay Kumar Mohanty, A Systematic Review on Recent Advancements in Deep Learning and Mathematical Modeling for Efficient Detection of Glioblastoma, 2024, 73, 0018-9456, 1, 10.1109/TIM.2024.3476544
    8. Muner. M. Abou Hasan, Seham. M. AL‐Mekhlafi, Fathala. A. Rihan, Hannah. A. Al‐Ali, Modeling Hybrid Crossover Dynamics of Immuno‐Chemotherapy and Gene Therapy: A Numerical Approach, 2025, 0170-4214, 10.1002/mma.10765
    9. Yuanhong Bi, Xiaoqi Zhang, Quansheng Liu, Deterministic and stochastic dynamic analysis of a Parkinson’s disease model, 2025, 195, 09600779, 116207, 10.1016/j.chaos.2025.116207
  • Reader Comments
  • © 2023 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(1948) PDF downloads(154) Cited by(9)

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog