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

Endemic coexistence and competition of virus variants under partial cross-immunity

  • In this study, we developed a mathematical framework, based on the SIR model, to study the dynamics of two competing virus variants with different characteristics of transmissibility, immune escape, and cross-immunity. The model includes variant-specific transmission and recovery rates and enables flexible parameterization of partial and waning cross-immunity. We conducted stability and bifurcation analyses and numerical simulations to explore the conditions of coexistence, dominance, and extinction of the variants, studying variations in epidemiological parameters that affect endemic prevalence and infection ratios. Our results indicated that transmission rates, levels of cross-immunity, and immunity waning rates are critical in determining disease outcomes, which influence variant prevalence and competitive dynamics. The sensitivity analysis provided the relative importance of these parameters and provided valuable insight into designing intervention strategies. This work contributes to furthering our understanding of multi-variant epidemic dynamics and lays the bedrock for tackling complex interactions involving arising virus variants, finding applications in real-world public health planning.

    Citation: Shirali Kadyrov, Farkhod Haydarov, Khudoyor Mamayusupov, Komil Mustayev. Endemic coexistence and competition of virus variants under partial cross-immunity[J]. Electronic Research Archive, 2025, 33(2): 1120-1143. doi: 10.3934/era.2025050

    Related Papers:

    [1] Xueyong Zhou, Xiangyun Shi . Stability analysis and backward bifurcation on an SEIQR epidemic model with nonlinear innate immunity. Electronic Research Archive, 2022, 30(9): 3481-3508. doi: 10.3934/era.2022178
    [2] Yike Lv, Xinzhu Meng . Evolution of infectious diseases induced by epidemic prevention publicity and interaction between heterogeneous strains. Electronic Research Archive, 2024, 32(8): 4858-4886. doi: 10.3934/era.2024223
    [3] Yujia Xiang, Yuqi Jiao, Xin Wang, Ruizhi Yang . Dynamics of a delayed diffusive predator-prey model with Allee effect and nonlocal competition in prey and hunting cooperation in predator. Electronic Research Archive, 2023, 31(4): 2120-2138. doi: 10.3934/era.2023109
    [4] Zimeng Lv, Jiahong Zeng, Yuting Ding, Xinyu Liu . Stability analysis of time-delayed SAIR model for duration of vaccine in the context of temporary immunity for COVID-19 situation. Electronic Research Archive, 2023, 31(2): 1004-1030. doi: 10.3934/era.2023050
    [5] Erhui Li, Qingshan Zhang . Global dynamics of an endemic disease model with vaccination: Analysis of the asymptomatic and symptomatic groups in complex networks. Electronic Research Archive, 2023, 31(10): 6481-6504. doi: 10.3934/era.2023328
    [6] Hui Cao, Mengmeng Han, Yunxiao Bai, Suxia Zhang . Hopf bifurcation of the age-structured SIRS model with the varying population sizes. Electronic Research Archive, 2022, 30(10): 3811-3824. doi: 10.3934/era.2022194
    [7] Ugo Avila-Ponce de León, Angel G. C. Pérez, Eric Avila-Vales . Modeling the SARS-CoV-2 sublineages XBB and BQ.1 in Mexico, considering multiple vaccinations, booster dose, waning immunity and cross-immunity. Electronic Research Archive, 2024, 32(2): 1082-1125. doi: 10.3934/era.2024053
    [8] Jorge Rebaza . On a model of COVID-19 dynamics. Electronic Research Archive, 2021, 29(2): 2129-2140. doi: 10.3934/era.2020108
    [9] Yuan Xue, Jinli Xu, Yuting Ding . Dynamics analysis of a diffusional immunosuppressive infection model with Beddington-DeAngelis functional response. Electronic Research Archive, 2023, 31(10): 6071-6088. doi: 10.3934/era.2023309
    [10] Weiyu Li, Hongyan Wang . Dynamics of a three-molecule autocatalytic Schnakenberg model with cross-diffusion: Turing patterns of spatially homogeneous Hopf bifurcating periodic solutions. Electronic Research Archive, 2023, 31(7): 4139-4154. doi: 10.3934/era.2023211
  • In this study, we developed a mathematical framework, based on the SIR model, to study the dynamics of two competing virus variants with different characteristics of transmissibility, immune escape, and cross-immunity. The model includes variant-specific transmission and recovery rates and enables flexible parameterization of partial and waning cross-immunity. We conducted stability and bifurcation analyses and numerical simulations to explore the conditions of coexistence, dominance, and extinction of the variants, studying variations in epidemiological parameters that affect endemic prevalence and infection ratios. Our results indicated that transmission rates, levels of cross-immunity, and immunity waning rates are critical in determining disease outcomes, which influence variant prevalence and competitive dynamics. The sensitivity analysis provided the relative importance of these parameters and provided valuable insight into designing intervention strategies. This work contributes to furthering our understanding of multi-variant epidemic dynamics and lays the bedrock for tackling complex interactions involving arising virus variants, finding applications in real-world public health planning.



    The ongoing threat of infectious diseases, particularly those capable of producing multiple viral variants, has underscored the need for advanced mathematical models to understand and predict the behavior of such diseases. New virus variants, as observed in the COVID-19 pandemic, are prone to exhibit significant differences in transmissibility, immune evasion capacity, and cross-immunity between variants. These differences are difficult for public health and epidemic modeling, considering that understanding the interaction of multiple virus variants in a population is essential in forecasting the trajectory of an epidemic and guiding intervention measures. Mathematical modeling has also been applied to investigate co-infections and the dynamics of their interactions, e.g., the dynamics of interaction between HTLV-2 and HIV-1 [1]. Here, an attempt is made to contribute towards such knowledge by developing and studying an SIR (susceptible-infected-recovered) model framework incorporating two competing virus variants with distinct epidemiological characteristics.

    Compartmental models like the SIR model, developed by Kermack and McKendrick [2], have been fundamental to epidemiology in dividing populations into susceptible, infected, and recovered compartments for simpler modeling of disease dynamics. Though effective in handling single-strain epidemics, standard SIR models are not flexible enough to capture complexity from interacting variants, hence making extensions for incorporating multi-variant dynamics a source of worry for scientists. The initial development of multi-strain or multi-variant models was necessitated by the reality that most infectious diseases, such as influenza and dengue fever, occur in more than one serotype or strain with varying epidemiological features. Multi-variant models attempt to capture these processes by introducing unique compartments or sub-compartments for each variant, each with varying transmission rates, recovery rates, and potential interactions, such as partial cross-immunity between variants. Initial pioneering work by Andreasen et al. [3] on multi-strain influenza models entailed competing strains coexisting in the presence of cross-immunity, and they found that transmissibility and degree of cross-immunity have a significant role in determining if variant strains can coexist or if one strain comes to dominate the other.

    Additional studies, such as those by Kamo and Sasaki [4], expanded these models by introducing seasonal forcing and analyzing how cross-immunity can lead to synchronized and chaotic oscillations in multi-strain epidemics. Garba and Gumel [5] further developed the understanding of influenza transmission dynamics by examining the effects of cross-immunity and backward bifurcation, underscoring the complex nature of multi-strain interactions. Similarly, Sneppen et al. [6] proposed a simplified model in which hosts can transmit only the most recent infection, creating a "spreading of immunity" effect that captures the diversity and duration of strain coexistence within a population.

    New research has developed these early tenets to include immune escape features and viral mutations, particularly for quickly evolving conditions like COVID-19. Mathematical models incorporating partial cross-immunity and asymmetric temporary durations of immunity have shown that minor differences in such factors can yield significant competitive disparity between strains [7,8]. For example, Ogura and Preciado's adaptive susceptible-infected-susceptible model demonstrates how individuals can sever connections to infected nodes to avoid disease spread, offering a new network-based epidemic control view [9]. Otunga [10] further developed this approach by modeling COVID-19 infections with focus on the Delta and Omicron variants, investigating how vaccination and recovery dynamics influence variant spread and control. Burbano Lombana et al. [11] employed a new approach by incorporating human behavior and memory impacts in models of simultaneous epidemic strains, which is important for the impact of interventions on multiple strains during COVID-19 epidemics. Olumoyin and Khaliq [12] employed a data-driven deep learning model for COVID-19 variant dynamics, which adopted methods to improve short-term prediction based on time-varying transmission rates. Bessonov et al. [13] also contributed to this understanding by considering immune responses, i.e., cytotoxic T lymphocytes and neutralizing antibodies, and how they can lead to the emergence and competition of respiratory virus variants and how immune escape mechanisms are established. The models considered in [7] show that early COVID-19 variants were more prone to the dramatic takeover while newer strains can potentially co-exist.

    The addition of cross-immunity and reinfection to such models has developed a more nuanced understanding of multi-variant epidemics. Cross-immunity occurs when immunity to a single strain cross-protects against another, a particularly relevant idea for viruses that are extremely mutable and lead to variants that can escape immune responses to previous infections. Reich et al. [14] employed a novel framework to examine cross-immunity among dengue serotypes and showed that cross-protection duration and intensity were of key importance to disease incidence and emphasized long-term immunization. Ferguson et al. [15] and Andreasen [16] studied the impact of moderate levels of cross-immunity on interaction dynamics and concluded that variants can become dominant temporarily before being replaced by new strains with greater immune evasion capacity, a phenomenon seen in influenza where strains become dominant in successive seasons due to loss of immunity.

    Further work has extended our understanding on how cross-immunity contributes to multi-variant behavior. Sachak-Patwa et al. [17] demonstrated that the integration of cross-immunity into models of influenza significantly enhances epidemic prediction, with the potential for further progress in public health policy. Atienza-Diez and Seoane [18] explored different types of cross-immunity, including sterilizing and attenuating, and concluded that higher levels of cross-immunity led to less frequent, reduced outbreaks. This shift in epidemic patterns emphasizes the effect of cross-immunity on the herd immunity levels and danger of future outbreaks. Chung and Lui [19] employed cross-immunity in a two-strain influenza model to demonstrate that alterations in cross-immunity levels may have significant effects on system stability and periodicity of outbreaks. For studies of COVID-19 variants, Niu et al. [20] developed a compartmental model with cross-immunity and heterogeneity of transmissibility for investigating stability and competition dynamics of variants.

    A number of studies provide broader perspectives on pathogen dynamics within host and ecological systems. For instance, Ojosnegros et al. [21] investigated the competition-colonization trade-offs among viral strains, finding that shifts in virulence distribution support a stable coexistence of low-virulence strains, highlighting viral evolution's complex interplay. Seabloom et al. [22] reviewed ecological interactions in pathogen spread, emphasizing the impact of species composition and community structure on infection dynamics, particularly in co-infection and host-pathogen relationships. Gjini et al. [23] focused on pneumococcus serotypes and the role of competition in pathogen dynamics, highlighting how direct competition shapes strain coexistence and challenges for vaccination strategies. Ackleh et al. [24] also contributed a reaction-diffusion model addressing competitive exclusion and conditions for pathogen strain coexistence. Amador et al. [25] analyzed stochastic interactions between antibiotic-sensitive and antibiotic-resistant bacterial strains, applying an extreme values approach to understand epidemic severity in hospital settings. Finally, Jover et al. [26] investigated host-phage dynamics in bacterial communities, finding that trade-offs in infection networks can influence the evolutionary and community structure of microbial ecosystems.

    While some researchers have made substantial progress in studying the dynamics of multi-variant systems, key knowledge gaps remain about what ensures the persistence, proliferation, or disappearance of competing molecular variants. Current models often crudely parameterize cross-immunity, treating it as either complete or non-existent, while empirical data suggest that cross-immunity is only partial, and depends on factors such as genetic relatedness between variants and immune responses of the host. Third, several of these assume that transmissibility and recovery rates of competing variants are constant, ignoring evolutionary pressures that will favor the emergence of more transmissible or immune-evasive variants.

    We aim to bridge these gaps by introducing an SIR model that incorporates two virus variants with distinct transmission and recovery rates, alongside a flexible framework for variable cross-immunity. Our aim is to explain how these factors interact to determine the competitive dynamics between variants. Using stability analysis and numerical simulations, we determine the equilibria behavior of variants — coexistence, dominance, and eradication — in the population. The biological significance of the model lies in its ability to simulate the dynamics of dual-variant epidemics, providing insights into how factors like transmissibility, immune evasion, and partial immunity influence viral spread and informing public health strategies.

    The proposed model aims to simulate the dynamics of an epidemic caused by two simultaneously circulating variants of a virus, which differ in terms of transmissibility, immune evasion, and recovery rates. The model divides the population into five compartments: Susceptible (S), infected by variant 1 (I1), infected by variant 2 (I2), and recovered individuals from variants 1 and 2 (R1 and R2, respectively).

    We assume that both strains of the virus spread separately, but their interaction through partial cross-immunity complicates the dynamics of the infection. The recovered population is assumed to have partial immunity to the other strain, i.e., individuals recovered from one strain can still be infected by the other but with reduced susceptibility. This provides a form of immunity that diminishes the risk of severe outcomes of reinfection, but not the risk of further transmission. We also assume that death from disease is negligible for both infectious strains, so it is not modeled. This enables the transmission dynamics, immunity, and recovery dynamics to be highlighted.

    The model is constructed under several assumptions to simplify the complexities of real-world epidemic scenarios. First, we assume that the total population remains constant, with births (Λ) and natural death (μ) rates included in the system. The transmission of each variant depends on the contact rate between susceptible individuals and infected individuals, with different transmission rates for each variant (β1 and β2). Furthermore, the recovery rates for each variant differ (γ1 and γ2), as do the rates of immunity loss for those recovered from either variant (δ).

    The model, whose flow diagram is given in Figure 1, is represented by a system of ordinary differential equations (ODEs) that describe the time evolution of the compartments. These equations incorporate the interactions between the compartments, including the effects of partial immunity due to prior recovery, as well as the dynamics of infection and recovery for each variant. The differential equations are as follows:

    dSdt=Λβ1SI1β2SI2+δ1R1+δ2R2μS,dI1dt=β1SI1+α1β1R2I1γ1I1μI1,dI2dt=β2SI2+α2β2R1I2γ2I2μI2,dR1dt=γ1I1α2β2R1I2δ1R1μR1,dR2dt=γ2I2α1β1R2I1δ2R2μR2. (2.1)
    Figure 1.  Flow diagram of the proposed model.

    Here, the variables represent the population sizes at time t for each compartment: S(t) for susceptible individuals, I1(t), and I2(t) for individuals infected with variants 1 and 2, respectively, and R1(t) and R2(t) for those recovered from variants 1 and 2, respectively. The parameters govern the interactions between these compartments, including the transmission rates for each variant (β1 and β2), the recovery rates (γ1 and γ2), the degree of cross-immunity between the variants (α1 and α2), and the rate of immunity loss (δ).

    This model provides a framework for understanding the dynamics of dual-variant epidemics, incorporating key features such as partial immunity and the potential for reinfection, which are crucial for modeling real-world viral outbreaks where multiple variants of a pathogen are in circulation. Through numerical simulations of the system of ODEs, the model can offer insights into the spread of both variants, the impact of different intervention strategies, and the long-term dynamics of dual-variant epidemics.

    The analysis of equilibria and stability forms the basis of all that is known about the dynamics of mathematical models in epidemiology. By determining the steady states and examining their stability, we can comprehend the potential for disease occurrence, eradication, and persistence. In this section, we rigorously explore the equilibria of the model, starting with the disease-free equilibrium, where there is no infection, and progressing towards endemic conditions to encompass the coexistence or dominance of infectious agents. With the help of the next-generation matrix, Jacobian analysis, and bifurcation analysis, we aim to explore thresholds and conditions to switch between these states. To find the steady states, we need to solve the system of equations

    0=Λβ1SI1β2SI2+δ1R1+δ2R2μS,0=β1SI1+α1β1R2I1γ1I1μI1,0=β2SI2+α2β2R1I2γ2I2μI2,0=γ1I1μR1α2β2R1I2δ1R1,0=γ2I2μR2α1β1R2I1δ2R2. (3.1)

    We see that disease-free equilibrium (DFE) is (S,I1,I2,R1,R2)=(Λ/μ,0,0,0,0). To compute the basic reproduction number R0, we use the next-generation matrix method. The dynamics of the infectious compartments are given by the equation

    dXdt=FV,

    where X=[I1I2], F represents the new infections, and V represents the transitions out of the infectious compartments. These are given as

    F=[β1SI1+α1β1R2I1β2SI2+α2β2R1I2],V=[γ1I1+μI1γ2I2+μI2].

    The Jacobian matrices at the disease-free equilibrium (DFE) are

    F=[β1Λμ00β2Λμ],V=[γ1+μ00γ2+μ].

    The next-generation matrix is given by FV1, where

    FV1=[β1Λμ00β2Λμ][1γ1+μ001γ2+μ]=[β1Λμ(γ1+μ)00β2Λμ(γ2+μ)].

    The basic reproduction number R0 is the spectral radius of FV1, which in this case is the maximum of the diagonal entries:

    Theorem 1. The system has basic reproduction number

    R0=max(β1Λμ(γ1+μ),β2Λμ(γ2+μ))

    and DFE is locally stable if and only if R0<1.

    As R0 crosses the critical value of 1, the system undergoes a bifurcation, transitioning from a stable disease-free state to a scenario where one or both viruses can persist in the population. This bifurcation marks a change in the qualitative dynamics of the system, highlighting the pivotal role of R0 in determining the system's behavior. For convenience, we fix the following notation

    R(1)0=β1Λμ(γ1+μ) and R(2)0=β2Λμ(γ2+μ),

    so that R0=max(R(1)0,R(2)0).

    An endemic equilibrium occurs when at least one of the infectious compartments is non-zero.

    Due to the symmetry of the model with respect to the two virus components, we focus on one of the two cases for a single-virus equilibrium in the population. Specifically, we consider the endemic equilibrium where I1=0 and I20. The situation where I2=0 and I10 is analogous and can be treated in a similar manner.

    Lemma 1. If R(2)0>1, the system admits a single virus endemic equilibrium (SVEE) given by

    (S,I1,I2,R1,R2)=(γ2+μβ2,0,R(2)01β2γ2μ(1γ2δ2(γ2+μ)(μ+δ2)),0,γ2μ+δ2I2). (3.2)

    Proof. When I1=0 and I20 we have R1=0,S=γ2+μβ2 and the system of Eq (3.1) simplifies to

    0=Λβ2SI2μS+δ2R2,0=γ2I2μR2δ2R2.

    Solving the second for R2 we get R2=γ2μ+δ2I2 and substituting this expression for R2 into the first equation

    0=Λβ2SI2μS+δ2γ2μ+δ2I2.

    Solving for I2 we get the desired result.

    Lemma 2. The endemic equilibrium from the above lemma is locally asymptotically stable if and only if

    β1S+α1β1R2γ1μ<0.

    In particular, if R(1)0R(2)0 the endemic equilibrium is unstable.

    Proof. Linearizing the system (2.1) around the single virus endemic equilibrium (3.2) gives

    ˙x=Ax

    for

    A=[A11β1Sβ2Sδ1δ2β1I1A2200α1β1I1β2I20A33α2β2I200γ1α2β2R1A4400α1β1R2γ20A55],

    where

    A11=β1I1β2I2μ,A22=β1S+α1β1R2γ1μ,A33=β2S+α2β2R1γ2μ,A44=δ1μα2β2I2,A55=δ2μα1β1I1.

    Let us consider the linearization around the endemic equilibrium as above. In particular, I1=R1=0 and A33=0 and S,I2,R2 are as in Lemma 1, giving

    A=[A11β1Sβ2Sδ1δ20A22000β2I200α2β2I200γ10A4400α1β1R2γ20A55].

    We proceed by calculating the eigenvalues.

    We see that

    A22=β1S+α1β1R2γ1μ

    is an eigenvalue and the remaining eigenvalues are determined from

    A=[A11β2Sδ1δ2β2I20α2β2I2000A4400γ20A55].

    Another eigenvalue is A44=δμα2β2I2<0, leaving

    A=[A11β2Sδ2β2I2000γ2A55].

    To compute the eigenvalues of the matrix we get

    |AλI|=det[A11λβ2Sδ2β2I2λ00γ2A55λ]=(λ+μ)det[111β2I2λ00γ2A55λ]=(λ+μ)[γ2β2I2+(A55λ)(λβ2I2)].

    Noting A55=δ2μ we arrive at

    (λ+μ)(λ2+(δ2+μ+β2I2)λ+β2I2(γ2+δ2+μ))=0.

    It follows that the last three eigenvalues are negative. Except for A22=β1S+α1β1R2γ1μ all other eigenvalues have negative real part. Hence, the stability depends on A22, which proves the first assertion of the lemma. As for the second assertion, we note that when β1S+α1β1R2γ1μ<0, it follows that

    β1S+α1β1R2γ1μ=β1Λμ((γ2+μ)μβ2Λ(γ1+μ)μβ1Λ+α1R2μΛ)>β1Λμ(1R(2)01R(1)0).

    We conclude that the endemic equilibrium becomes unstable if R(1)0R(2)0.

    In this section, we study the endemic equilibrium when both viruses coexist, i.e., when I10 and I20. Under these conditions, the system of equations simplifies, enabling us to express key variables such as the susceptible population S in terms of a cubic equation. To simplify the formulas, we fix the following notations:

    A1:=γ1+μβ1,A2:=γ2+μβ2,a1:=α1γ2β2,a2:=α2γ1β1. (3.3)

    Proposition 1. If system (2.1) admits a coexistence equilibrium (I1,I20) then the number of susceptible S at the equilibrium satisfies the cubic polynomial

    c3S3+c2S2+c1S+c0=0,

    with coefficients

    c3=μα1α2(α1(1α2)+α2),c2=α1α2(μA1α1μA2α1μA1α2μA2α2+Λα1α2+μA1α1α2+μA2α1α2+A2α1δ1a1(μ+δ1)+A1α2δ2a2(μ+δ2)),c1=a1(A2α1α2(μ+δ1)+a2(α1(μα2+δ1)+α2δ2))α1α2(A2α1(Λα2+A2δ1)+A21α2δ2a2A1(μ+δ2)+A1(Λα1α2+A2(α1(μ+μα2+δ1)+α2(μ+δ2)))),c0=((a1a2A1A2α1α2)(α1(Λα2+A2δ1)+A1α2δ2)).

    Moreover, S,I1,I2 at the equilibrium must satisfy

    Λβ1I1+β2I2+μSΛμ1R0.

    It remains unclear whether the necessary conditions stated in the proposition are also sufficient. While it is straightforward to verify that the coefficients c0 and c3 are both positive, ensuring that at least one root of the cubic equation is negative, the possibility of the existence of two distinct coexistence equilibria is not established. This ambiguity warrants further investigation. Our numerical analysis in § 5 demonstrates the existence of a coexistence equilibrium under certain conditions. In addition to establishing existence, conducting a stability analysis is equally crucial for a comprehensive understanding. Again, our numerical analysis in § 5 shows that the existence of coexistence equilibrium does not necessarily imply asymptotic stability.

    Proof. When I10 and I20, the system of equations simplifies to

    0=Λβ1SI1β2SI2+δ1R1+δ2R2μS,0=β1S+α1β1R2γ1μ,0=β2S+α2β2R1γ2μ,0=γ1I1μR1α2β2R1I2δ1R1,0=γ2I2μR2α1β1R2I1δ2R2.

    Solving for R1 and R2 we obtain

    δ1R1+δ2R2=β1SI1+β2SI2+μSΛ, (3.4)
    α1R2=γ1+μβ1S, (3.5)
    α2R1=γ2+μβ2S, (3.6)
    α2R1=α2γ1I1α2β2I2+δ1+μ, (3.7)
    α1R2=α1γ2I2α1β1I1+δ2+μ. (3.8)

    Since R1,R20 we obtain the last assertion in the proposition. Letting x=β1SI1,y=β2SI2 and substituting for α1R2 and α2R1, we get

    δ1α2(A2S)+δ2α1(A1S)=x+y+μSΛ, (3.9)
    a1yα1x+(δ2+μ)S=A1S, (3.10)
    a2xα2y+(δ1+μ)S=A2S. (3.11)

    This gives

    a1y=(A1S)(α1x+(δ2+μ)S)=a1(δ1α2(A2S)+δ2α1(A1S)xμS+Λ),a2x=(A2S)(α2y+(δ1+μ)S)=a2(δ1α2(A2S)+δ2α1(A1S)yμS+Λ).

    Solving for x and y:

    x=a1δ1α2(A2S)+a1δ2α1(A1S)a1μS+a1Λ(A1S)(δ2+μ)S(A1S)α1+a1,y=a2δ1α2(A2S)+a2δ2α1(A1S)a2μS+a2Λ(A2S)(δ1+μ)S(A2S)α2+a2.

    Substituting this into

    δ1α2(A2S)+δ2α1(A1S)=x+y+μSΛ

    and solving for S yields cubic equation:

    δ1α2(A2S)+δ2α1(A1S)=a1δ1α2(A2S)+a1δ2α1(A1S)a1μS+a1Λ(A1S)(δ2+μ)S(A1S)α1+a1+a2δ1α2(A2S)+a2δ2α1(A1S)a2μS+a2Λ(A2S)(δ1+μ)S(A2S)α2+a2+μSΛ.

    Simplifying this yields the desired cubic polynomial.

    In Theorem 1, we establish the local stability of the disease-free equilibrium (DFE). In this section, we establish the global stability of DFE. To achieve this, we construct a suitable Lyapunov function and analyze the system dynamics within the positively invariant set Ω. The set Ω ensures that all solutions remain biologically feasible, and using standard techniques, we demonstrate that when R0<1, the DFE is asymptotically stable within Ω. This result highlights the conditions under which the disease can be eradicated from the population, building on approaches seen in similar analyses [27].

    We define

    Ω:={(S,I1,I2,R1,R2)R50S+I1+I2+R1+R2Λ/μ}. (4.1)

    Thus, we have the following.

    Lemma 3. The set Ω given in (4.1) is positively invariant, that is any solution to (2.1) with initial values in Ω stays in Ω for t0.

    This lemma follows from standard arguments, as those in [28].

    Theorem 2. When R0<1, the DFE (S,I1,I2,R1,R2)=(Λ/μ,0,0,0,0) is asymptotically stable on Ω.

    Proof. Let us consider the Lyapunov function given by

    V(t)=a(ΛμS(t))+1γ1+μI1(t)+1γ2+μI2(t)+aR1(t)+aR2(t),

    where the coefficients a,b>0 are sufficiently small to be determined later. We now show that V is positive definitive and dVdt is negative definite on Ω.

    We note that V>0 on Ω except at DFE. Hence, it is positive definite. To show negative definiteness, let us write V=G+F where

    G=a(ΛμS)+b(R1+R2) and F=1γ1+μI1+1γ2+μI2.

    We have

    dFdt=1γ1+μdI1dt+1γ2+μdI2dt=I1(β2(S+α1R2)γ1+μ1)+I2(β2(S+α2R1)γ2+μ1).

    Since α1,α21 we see that S+α1R2S+R2Λ/μ on Ω. Similarly, S+α2R1Λ/μ. Thus,

    dFdtI1(β1Λ(γ1+μ)μ1)+I2(β2Λ(γ2+μ)μ1)(I1+I2)(R01). (4.2)

    Next, we consider the time derivative of G:

    dGdt=a(Λβ1SI1β2SI2+δ1R1+δ2R2μS)+a(γ1I1α2β2R1I2δ1R1μR1+γ2I2α1β1R2I1δ2R2μR2).

    Using ΛμS0 on Ω and denoting by H the negative terms

    H(S,R1,R2)=a(ΛμS+(2δ1+μ)R1+(2δ2+μ)R2), (4.3)

    we get

    dGdt=aβ1SI1+aβ2SI2+aγ1I1+aγ2I2+H(S,R1,R2)aI1(β1Λμ+γ1)+aI2(β2Λμ+γ2)+H(S,R1,R2).

    We are given that R0<1. We may pick a>0 sufficiently small so that

    amax(β1Λμ+γ1,β2Λμ+γ2)1R02.

    With this choice of a we have

    dGdt(I1+I2)1R02+H(S,R1,R2). (4.4)

    Combining (4.2) and (4.4) we arrive at

    dVdt=dFdt+dGdt(I1+I2)(R01)+(I1+I2)1R02+H(S,R1,R2)=(I1+I2)R012+H(S,R1,R2).

    Since R0<1, we see from (4.3) that dV/dt<0 on Ω except at DFE. Hence, it follows from Lyapunov stability theorems, see e.g., [29, Theorem 5.2] and [30, Theorem 4.1], that DFE is asymptotically stable on Ω.

    For calibration, we use dynamics of Delta and Omicron variants of COVID-19. The competitive dynamics of virus variants, such as Delta and Omicron, highlight the pivotal role of partial cross-immunity in shaping epidemiological outcomes [31]. The Delta variant, first identified in India in late 2020, became the dominant strain worldwide by mid-2021 due to its high transmissibility and virulence. Omicron, detected in South Africa and Botswana in November 2021, rapidly outpaced the Delta strain to become the dominant variant by December 2021 in many regions, including the United States. This rapid displacement was attributed to the reproduction number of Omicron, which was approximately 3.5 times higher than that of Delta, coupled with enhanced immune evasion capabilities.

    Table 1 presents the parameter values used for model calibration. Unless otherwise specified, we use these values in our numerical analysis. Using these values, the basic reproduction numbers for the Delta and Omicron variants are calculated as R(1)0=β1Λμ(γ1+μ)=3.80 and R(2)0=β2Λμ(γ2+μ)=8.29, respectively. These values fall within the ranges previously reported in the literature (see, e.g., [35,36]). We note that letting μ=Λ we assume the constant population model with a total population density equal to 1.

    Table 1.  Parameter values for model calibration.
    Parameter Symbol Value Source
    Natural death rate μ 3.9×105
    Recruitment rate Λ 3.9×105
    Transmission rate (Delta) β1 0.3317 [31]
    Transmission rate (Omicron) β2 0.9951 [31]
    Recovery rate (Delta) γ1 0.1 [32]
    Recovery rate (Omicron) γ2 0.12 [32]
    Cross-immunity (Delta to Omicron) α2 0.3 [33]
    Cross-immunity (Omicron to Delta) α1 0.25 [33]
    Natural immunity waning rate δ1,δ2 0.0037 [34]

     | Show Table
    DownLoad: CSV

    Table 2 outlines the initial population values used in the numerical simulations. These baseline conditions provide a consistent framework for analyzing the infection dynamics and steady-state behavior across the scenarios presented in this study.

    Table 2.  Initial population values.
    Variable Initial values
    I1(0) 0.100
    I2(0) 0.001
    R1(0) 0.200
    R2(0) 0.000
    S(0) 0.699

     | Show Table
    DownLoad: CSV

    Figure 2 illustrates the dynamics of the model through 100 simulations, where the parameters are randomly selected within a ±20% range as specified in Table 1.

    Figure 2.  Model dynamics for parameter values.

    In numerical analysis, we are interested in studying the endemic prevalence (I1+I2)(t) and the infection ratio (I1/(I1+I2))(t), particularly their final values.

    While Lemma 2 provides criteria for local stability, it does not offer insights into global dynamics. Therefore, we conduct numerical simulations to better illustrate the global dynamics, the prevalence, and their relationship to the stability of the single-virus dominance equilibrium.

    In Figure 3, we present three representative examples of the infection dynamics over 100,000 days for different transmission rates (β1 and β2). These figures illustrate the behavior of the two-virus system, highlighting how variations in transmission parameters influence the temporal evolution of the infected populations (I1(t) and I2(t)). The results show distinct patterns of dominance, coexistence, or decline of the viruses, providing a visual foundation for the numerical analysis summarized in the following table. Looking at the last 1000 days of the simulation, we observe periodic behavior, indicating that the infection may not necessarily converge to any of the equilibria.

    Figure 3.  Infection dynamics over 100,000 days for various transmission rates.

    Table 3 presents numerical results for varying values of β1 and β2. The table includes the basic reproduction numbers R0, R(1)0, and R(2)0, as well as the corresponding endemic prevalences and infection ratios after 100k days. The values of I1 and I2 represent the endemic equilibria for the single-virus cases where I1=0 and I2=0, respectively. In addition, the coexistence equilibrium (I1c,I2c) is shown, reflecting the steady-state values of both infected populations when both viruses coexist. The notation, (u) and (s) indicate whether the equilibrium is locally asymptotically unstable or stable, respectively. We note that when the infection ratio (IR) lies strictly between 0 and 1, both virus variants coexist in the population. From the table, it is evident that as R0 decreases, the infection ratio (IR) approaches 1.0, indicating a higher prevalence of one virus over the other or a dominance of the coexistence equilibrium. The bifurcation of the system occurs as R0 passes critical thresholds, transitioning between stable and unstable equilibria, with the corresponding changes in the infection dynamics. The entries marked as "DNE" (Does Not Exist) suggest that certain conditions lead to a scenario where no valid equilibrium can be reached. These cases typically occur when the basic reproduction number R0 is too low to sustain either virus at endemic levels. Endemic Prevalence (EP) represents the steady-state infection levels of each variant in the population. It reflects the long-term dynamics of the virus after the system has reached equilibrium, providing insights into how the infection stabilizes over time under different conditions.

    Table 3.  Numerical results for varying β1,β2.
    R0 R(1)0 R(2)0 I1 I2 (I1c,I2c) EP IR
    9.65 0.11 9.65 DNE 0.0033 (s) (0.023, 0.001) (s) 0.027 0.00
    7.73 3.86 7.73 0.0027 (u) 0.0032 (s) (0.021, 0.034) (u) 0.054 0.38
    5.21 3.45 5.21 0.0026 (u) 0.0029 (s) (0.017, 0.027) (u) 0.045 0.39
    3.92 2.43 3.92 0.0021 (u) 0.0027 (s) (0.004, 0.023) (u) 0.026 0.15
    2.83 2.83 1.91 0.0023 (s) 0.0017 (u) (0.023, 0.001) (u) 0.024 0.94
    2.64 0.65 2.64 DNE 0.0023 (s) DNE 0.019 0.00
    1.77 1.77 1.23 0.0016 (s) 0.0007 (u) DNE 0.016 1.00
    1.51 1.51 1.29 0.0012 (s) 0.0008 (u) DNE 0.012 1.00
    1.23 1.23 0.96 0.0007 (s) DNE DNE 0.007 1.00
    0.29 0.29 0.14 DNE DNE DNE 0.000 DNE

     | Show Table
    DownLoad: CSV

    For instance, when R0=9.65, I1 does not exist (denoted by DNE), while I2 has a small value, and the coexistence equilibrium is nonzero. This suggests that the system tends to focus on the second virus as the more dominant strain when the first virus does not establish a steady state. As R0 decreases to values near 1, the table indicates that the system shifts toward a state where both viruses coexist with low prevalences, resulting in a low infection ratio (IR), and eventually both viruses fail to persist at all when R0 falls below 1.

    In the next, we consider prevalence and infection ratio for varying cross immunities α1 and α2 (Figure 4). This is done by considering the 10,000 day iteration of the system and the variablees are calculated by taking the average of the last 100 days from simulations. The left panel shows that higher cross-immunity tends to reduce total prevalence, while the right panel indicates that infection ratio dynamics are more sensitive to variations in α1, suggesting potential asymmetric effects of cross-immunity on variant competition.

    Figure 4.  Contour plots for for varying α1 and α2 at the end of 10,000 days.

    To enrich the practical applicability of our results, we note the implications for public health strategies of the multi-variant nature of the epidemics. Indeed, as our analysis points out, multiple endemic equilibria are possible depending on certain factors like cross-immunities and transmission rates. If reaching the DFE is not immediate, strategies might be developed that would drive the dynamics toward the equilibrium with the lowest magnitude of prevalence. This could be achieved by interventions targeting a selective reduction in the dominance of variants of higher transmissibility or immune evasion that, in the end, reduce overall disease burden.

    We analyze the sensitivity of two key outcomes in a multi-strain infectious disease model—total prevalence and infection ratio—to various epidemiological parameters. Sensitivity analysis is a crucial tool in understanding how the variability in model outputs can be attributed to different input variables [37]. Using Latin Hypercube Sampling (LHS), 2000 parameter sets are generated for 8 parameters (β1,β2,γ1,γ2,α1,α2,δ1,δ2) across ranges spanning from 0.5 to 1.5 times the base values, as provided in Table 1. The model, solved numerically using ODEs, simulates the dynamics of susceptible, infected, and recovered populations, calculating average steady-state values of the two dependent variables.

    Sensitivity analysis is conducted by calculating the Partial Rank Correlation Coefficient (PRCC) using Spearman's rank correlation method, and the results are visualized in Figure 5. The bar charts show the strength and direction of correlations for each parameter, with statistically significant results (p<0.05) marked by red asterisks. Parameters such as transmission rates (β1,β2) strongly influence Total Infected, while interaction terms (α1,α2) significantly impact infection ratio. These findings highlight the most influential parameters, providing insights into disease dynamics and intervention strategies.

    Figure 5.  Partial rank correlation coefficients (PRCC) derived from 2000 simulations for (a) total prevalence and (b) infection ratio, indicating the relative sensitivity of parameters. Red asterisks denote statistically significant correlations (p<0.05).

    Additionally, convergence of the PRCC results was assessed to ensure that the sensitivity analysis stabilized with increasing sample size. Figure 6 displays the convergence plots for total infected (a) and infection ratio (b), demonstrating the stabilization of partial rank correlation coefficients as the number of simulations increases. These plots confirm the reliability of the sensitivity analysis, with results becoming more consistent as the simulation count approaches 1000.

    Figure 6.  Convergence of partial rank correlation coefficients (PRCC) for the model outcomes with 1000 simulations. (a) Total infected population and (b) infection ratio illustrate the stabilization of sensitivity analysis as the number of simulations grows.

    For completeness, scatter plots illustrating the relationships between parameters and dependent variables are provided in Figures 7 and 8. These plots offer a detailed visual representation of how each epidemiological parameter influences the model outcomes—endemic prevalence and infection ratio—across simulation scenarios. By plotting the parameter values against the corresponding model outputs, these scatter plots enable a clearer understanding of the magnitude and direction of the parameter effects, complementing the sensitivity analysis presented earlier.

    Figure 7.  Scatter plots showing the relationship between model parameters and endemic prevalence. Each plot illustrates how a specific parameter (β1,β2,γ1,γ2,α1,α2,δ1,δ2) influences the endemic prevalence across different simulation scenarios.
    Figure 8.  Scatter plots depicting the correlation between model parameters and infection ratio. These plots visualize the impact of parameters (β1,β2,γ1,γ2,α1,α2,δ1,δ2) on the infection ratio, providing insights into their relative contribution to the disease dynamics.

    Figure 5(a) demonstrates the sensitivity of total prevalence to various epidemiological parameters using Partial Rank Correlation Coefficients (PRCC). The transmission rates (β1 and β2) exhibit the strongest positive correlations, indicating that higher transmission rates significantly increase the total prevalence of infection. Recovery rates (γ1 and γ2) are negatively correlated with prevalence, underscoring that faster recovery leads to reduced infection levels. Waning immunity rates (δ1 and δ2) show weaker positive correlations, suggesting that quicker loss of immunity slightly amplifies the spread by increasing susceptibility to reinfection. These findings highlight the pivotal role of transmission and recovery dynamics in shaping the overall disease burden.

    Figure 5(b) reveals the sensitivity of the infection ratio to key model parameters, focusing on the distribution of infections between the two virus variants. Cross-immunity coefficients (α1 and α2) emerge as the most influential factors, with significant correlations that reflect their role in modulating variant competition and dominance. The transmission rates (β1 and β2) also influence the infection ratio, demonstrating how disparities in transmissibility between variants can shift the balance of prevalence.

    These results highlight the importance of sensitivity analysis in designing public health interventions that control multi-strain infectious diseases. Key parameters such as transmission rates β1,β2, cross-immunity coefficients α1,α2, and immunity waning rates δ1,δ2 strongly influence the robustness of model outcomes. It is seen from the above discussions that parameters, such as the transmission rates β1, β2 and the cross-immunity coefficients α1, α2, have a profound effect on both the total prevalence and infection ratios to necessitate site-specific interventions adaptive to the change in epidemic conditions. Besides, the minor but significant estimate of immunity waning rates δ1,δ2 points out the need for real-time data application to revise strategies for long-term effective management of the epidemic. It is hoped that researchers may seek the development of adaptive measures that could better achieve an optimized response with resources based on continued shifts in disease dynamics.

    The sensitivity analysis highlights that the dynamics of multivariant epidemics are deeply influenced by changes in key parameters, such as transmission rates and cross-immunity. These findings have contributed to informing public health strategies since they may be used to devise interventions for particular epidemic scenarios. For example, targeting the most influential parameters-such as enhancing cross-immunity or adjusting vaccination strategies-may help mitigate the spread of more transmissible or immune-evasive variants. The discussion on adaptive intervention strategies, active mechanisms of feedback, and control-oriented implications might provide greater depth. Integration of adaptive measures from real-time epidemic data is key to optimally responding to fluctuating transmission rates and evolving variants. This enables dynamic adjustments that could work towards effective control of the disease burden. Understanding these factors-precisely how those factors influence the long-term course of the epidemics-will help stakeholders build up adaptive responses that are more effective by optimizing resource allocation and timely intervention.

    In this paper, we provide a comprehensive analysis of the dynamics of two competing virus variants within an SIR framework, emphasizing the role of partial cross-immunity and other epidemiological factors. By exploring the sensitivity of key outcomes, such as total prevalence and infection ratio, to model parameters, we indetify critical drivers of disease dynamics. Transmission rates emerge as the most significant contributors to total prevalence, while cross-immunity levels play a pivotal role in shaping the infection ratio and the competitive interplay between variants.

    Our findings underscore the complexity of multi-variant epidemics, highlighting the intricate balance between factors like recovery rates, waning immunity, and cross-immunity, which are invaluable for public health planning. These insights hold great practical implications for designing interventions to effectively reduce disease burden and manage variant competition. For example, actions to enhance cross-immunity or act on enhanced transmissibility of new variants can stem the tide of more virulent or immune-escape strains. Enhanced cross-immunity through vaccination, enhancing monitoring and surveillance infrastructure for tracking variant spread, and creating effective communications framing vaccination and cross-immunity are essential steps to managing variant threats. Furthermore, adaptive response strategies tailored to various epidemiological contexts will enable more effective interventions against newly emerging variants. Together, these findings provide a more nuanced understanding of viral competition dynamics to help stakeholders better prepare for and respond to multi-variant infectious disease challenges.

    While the mathematical model yields robust theoretical predictions, their implications are also pertinent for rapidly evolving diseases in real-world settings. Furthermore, researchers may examine how various population structures, behavioral responses, and intervention strategies influence these dynamics. Overall, this study highlights the necessity of a refined understanding of variant interactions for informing adaptive and evidence-based public health interventions.

    Shirali Kadyrov: Conceptualization, Formal analysis, Investigation, Methodology, Writing – original draft, Writing – review & editing. Farkhod Khaydarov: Formal analysis, Investigation, Writing – review & editing Khudoyor Mamayusupov: Formal analysis, Investigation, Writing – review & editing Komil Mustayev: Formal analysis, Investigation, Software, Visualization, Writing – original draft, Writing – review & editing.

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

    The first author acknowledges the support of a grant from the Ministry of Science and Higher Education of the Republic of Kazakhstan within the framework of the project AP19676669.

    Authors declare no conflict of interest.



    [1] A. Elaiw, E. Almohaimeed, A. Hobiny, Modeling the co-infection of HTLV-2 and HIV-1 in vivo, Electron. Res. Arch., 32 (2024), 6032–6071. https://doi.org/10.3934/era.2024280 doi: 10.3934/era.2024280
    [2] W. O. Kermack, A. G. McKendrick, A contribution to the mathematical theory of epidemics, Proc. R. Soc. London, Ser. A, 115 (1927), 700–721. https://doi.org/10.1098/rspa.1927.0118 doi: 10.1098/rspa.1927.0118
    [3] V. Andreasen, J. Lin, S. A. Levin, The dynamics of cocirculating influenza strains conferring partial cross-immunity, J. Math. Biol., 35 (1997), 825–842. https://doi.org/10.1007/s002850050079 doi: 10.1007/s002850050079
    [4] M. Kamo, A. Sasaki, The effect of cross-immunity and seasonal forcing in a multi-strain epidemic model, Physica D: Nonlinear Phenom., 165 (2002), 228–241. https://doi.org/10.1016/S0167-2789(02)00389-5 doi: 10.1016/S0167-2789(02)00389-5
    [5] S. M. Garba, A. B. Gumel, Effect of cross-immunity on the transmission dynamics of two strains of dengue, Int. J. Comput. Math., 87 (2010), 2361–2384. https://doi.org/10.1080/00207160802660608 doi: 10.1080/00207160802660608
    [6] K. Sneppen, A. Trusina, M. H. Jensen, S. Bornholdt, A minimal model for multiple epidemics and immunity spreading, PloS One, 5 (2010), e13326. https://doi.org/10.1371/journal.pone.0013326 doi: 10.1371/journal.pone.0013326
    [7] M. D. Johnston, B. Pell, D. Rubel, A two-strain model of infectious disease spread with asymmetric temporary immunity periods and partial cross-immunity, Math. Biosci. Eng., 20 (2023), 16083–16113, https://doi.org/10.3934/mbe.2023718 doi: 10.3934/mbe.2023718
    [8] B. Pell, S. Brozak, T. Phan, F. Wu, Y. Kuang, The emergence of a virus variant: Dynamics of a competition model with cross-immunity time-delay validated by wastewater surveillance data for COVID-19, J. Math. Biol., 86 (2023). https://doi.org/10.1007/s00285-023-01900-0
    [9] M. Ogura, V. M. Preciado, Epidemic processes over adaptive state-dependent networks, Phys. Rev. E, 93 (2016), 062316. https://doi.org/10.1103/PhysRevE.93.062316 doi: 10.1103/PhysRevE.93.062316
    [10] O. M. Otunuga, Analysis of multi-strain infection of vaccinated and recovered population through epidemic model: Application to COVID-19, PloS One, 17 (2022), e0271446. https://doi.org/10.1371/journal.pone.0271446 doi: 10.1371/journal.pone.0271446
    [11] D. A. B. Lombana, L. Zino, S. Butail, E. Caroppo, Z. P. Jiang, A. Rizzo, et al., Activity-driven network modeling and control of the spread of two concurrent epidemic strains, Appl. Network Sci., 7 (2022), 66. https://doi.org/10.1007/s41109-022-00507-6 doi: 10.1007/s41109-022-00507-6
    [12] K. Olumoyin, A. Khaliq, Multi-variant COVID-19 model with heterogeneous transmission rates using deep neural networks, preprint, arXiv: 2205.06834.
    [13] N. Bessonov, D. Neverova, V. Popov, V. Volpert, Emergence and competition of virus variants in respiratory viral infections, Front. Immunol., 13 (2023), 945228. https://doi.org/10.3389/fimmu.2022.945228 doi: 10.3389/fimmu.2022.945228
    [14] N. G. Reich, S. Shrestha, A. A. King, P. Rohani, J. Lessler, S. Kalayanarooj, et al., Interactions between serotypes of dengue highlight epidemiological impact of cross-immunity, J. R. Soc. Interface, 10 (2013), 20130414. https://doi.org/10.1098/rsif.2013.0414 doi: 10.1098/rsif.2013.0414
    [15] N. M. Ferguson, A. P. Galvani, R. M. Bush, Ecological and immunological determinants of influenza evolution, Nature, 422 (2003), 428–433. https://doi.org/10.1038/nature01509 doi: 10.1038/nature01509
    [16] V. Andreasen, Epidemics in competition: Partial cross-immunity, Bull. Math. Biol., 80 (2018), 2957–2977. https://doi.org/10.1007/s11538-018-0495-2 doi: 10.1007/s11538-018-0495-2
    [17] R. Sachak-Patwa, H. M. Byrne, R. N. Thompson, Accounting for cross-immunity can improve forecast accuracy during influenza epidemics, Epidemics, 34 (2021), 100432. https://doi.org/10.1016/j.epidem.2020.100432 doi: 10.1016/j.epidem.2020.100432
    [18] I. Atienza-Diez, L. F. Seoane, Long-and short-term effects of cross-immunity in epidemic dynamics, Chaos, Solitons & Fractals, 174 (2023), 113800. https://doi.org/10.1016/j.chaos.2023.113800 doi: 10.1016/j.chaos.2023.113800
    [19] K. Chung, R. Lui, Dynamics of two-strain influenza model with cross-immunity and no quarantine class, J. Math. Biol., 73 (2016), 1467–1489. https://doi.org/10.1007/s00285-016-1000-x doi: 10.1007/s00285-016-1000-x
    [20] R. Niu, Y. C. Chan, S. Liu, E. W. Wong, M. A. van Wyk, Stability analysis of an epidemic model with two competing variants and cross-infections, preprint. https://doi.org/10.21203/rs.3.rs-3264948/v1
    [21] S. Ojosnegros, E. Delgado-Eckert, N. Beerenwinkel, Competition–colonization trade-off promotes coexistence of low-virulence viral strains, J. R. Soc. Interface, 9 (2012), 2244–2254. https://doi.org/10.1098/rsif.2012.0160 doi: 10.1098/rsif.2012.0160
    [22] E. W. Seabloom, E. T. Borer, K. Gross, A. E. Kendig, C. Lacroix, C. E. Mitchell, et al., The community ecology of pathogens: Coinfection, coexistence and community composition, Ecol. Lett., 18 (2015), 401–415. https://doi.org/10.1111/ele.12418 doi: 10.1111/ele.12418
    [23] E. Gjini, C. Valente, R. Sa-Leao, M. G. M. Gomes, How direct competition shapes coexistence and vaccine effects in multi-strain pathogen systems, J. Theor. Biol., 388 (2016), 50–60. https://doi.org/10.1016/j.jtbi.2015.09.031 doi: 10.1016/j.jtbi.2015.09.031
    [24] A. S. Ackleh, K. Deng, Y. Wu, Competitive exclusion and coexistence in a two-strain pathogen model with diffusion, Math. Biosci. Eng., 13 (2015), 1–18. https://doi.org/10.3934/mbe.2016.13.1 doi: 10.3934/mbe.2016.13.1
    [25] J. Amador, D. Armesto, A. Gómez-Corral, Extreme values in sir epidemic models with two strains and cross-immunity, Math. Biosci. Eng., 16 (2019), 1992–2022. https://doi.org/10.3934/mbe.2019098 doi: 10.3934/mbe.2019098
    [26] L. F. Jover, M. H. Cortez, J. S. Weitz, Mechanisms of multi-strain coexistence in host–phage systems with nested infection networks, J. Theor. Biol., 332 (2013), 65–77. https://doi.org/10.1016/j.jtbi.2013.04.011 doi: 10.1016/j.jtbi.2013.04.011
    [27] R. Rifhat, K. Wang, L. Wang, T. Zeng, Z. Teng, Global stability of multi-group seiqr epidemic models with stochastic perturbation in computer network, Electron. Res. Arch., 31 (2023), 4155–4184. https://doi.org/10.3934/era.2023212 doi: 10.3934/era.2023212
    [28] H. G. Anderson, G. P. Takacs, D. C. Harris, Y. Kuang, J. K. Harrison, T. L. Stepien, Global stability and parameter analysis reinforce therapeutic targets of PD-l1-PD-1 and MDSCs for glioblastoma, J. Math. Biol., 88 (2024), 10. https://doi.org/10.1007/s00285-023-02027-y doi: 10.1007/s00285-023-02027-y
    [29] F. Brauer, J. A. Nohel, The Qualitative Theory of Ordinary Differential Equations: An Introduction, Courier Corporation, 1989.
    [30] A. Lajmanovich, J. A. Yorke, A deterministic model for gonorrhea in a nonhomogeneous population, Math. Biosci., 28 (1976), 221–236. https://doi.org/10.1016/0025-5564(76)90125-5 doi: 10.1016/0025-5564(76)90125-5
    [31] C. N. Ngonghala, H. B. Taboe, S. Safdar, A. B. Gumel, Unraveling the dynamics of the omicron and delta variants of the 2019 coronavirus in the presence of vaccination, mask usage, and antiviral treatment, Appl. Math. Modell., 114 (2023), 447–465. https://doi.org/10.1016/j.apm.2022.09.017 doi: 10.1016/j.apm.2022.09.017
    [32] C. Menni, A. M. Valdes, L. Polidori, M. Antonelli, S. Penamakuri, A. Nogal, et al., Symptom prevalence, duration, and risk of hospital admission in individuals infected with SARS-COV-2 during periods of omicron and delta variant dominance: A prospective observational study from the zoe covid study, The Lancet, 399 (2022), 1618–1624. https://doi.org/10.1016/S0140-6736(22)00327-0 doi: 10.1016/S0140-6736(22)00327-0
    [33] S. Liossi, E. Tsiambas, S. Maipas, E. Papageorgiou, A. Lazaris, N. Kavantzas, Mathematical modeling for delta and omicron variant of SARS-COV-2 transmission dynamics in greece, Infect. Dis. Modell., 8 (2023), 794–805. https://doi.org/10.1016/j.idm.2023.07.002 doi: 10.1016/j.idm.2023.07.002
    [34] C. Cassata, How long does immunity last after COVID-19? What we know, Healthline, 2021 (2021).
    [35] Y. Liu, J. Rocklöv, The reproductive number of the delta variant of SARS-COV-2 is far higher compared to the ancestral SARS-COV-2 virus, J. Travel Med., 28 (2021), taab124. https://doi.org/10.1093/jtm/taab124 doi: 10.1093/jtm/taab124
    [36] Y. Liu, J. Rocklöv, The effective reproductive number of the omicron variant of SARS-COV-2 is several times relative to delta, J. Travel Med., 29 (2022), taac037. https://doi.org/10.1093/jtm/taac037 doi: 10.1093/jtm/taac037
    [37] L. Hanum, D. Ertiningsih, N. Susyanto, Sensitivity analysis unveils the interplay of drug-sensitive and drug-resistant glioma cells: Implications of chemotherapy and anti-angiogenic therapy, Electron. Res. Arch., 32 (2024), 72–89. https://doi.org/10.3934/era.2024004 doi: 10.3934/era.2024004
  • Reader Comments
  • © 2025 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(436) PDF downloads(34) Cited by(0)

Figures and Tables

Figures(8)  /  Tables(3)

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog