
Electronic Medical Record (EMR) is the data basis of intelligent diagnosis. The diagnosis results of an EMR are multi-disease, including normal diagnosis, pathological diagnosis and complications, so intelligent diagnosis can be treated as multi-label classification problem. The distribution of diagnostic results in EMRs is imbalanced. And the diagnostic results in one EMR have a high coupling degree. The traditional rebalancing methods does not function effectively on highly coupled imbalanced datasets. This paper proposes Double Decoupled Network (DDN) based intelligent diagnosis model, which decouples representation learning and classifier learning. In the representation learning stage, Convolutional Neural Networks (CNN) is used to learn the original features of the data. In the classifier learning stage, a Decoupled and Rebalancing highly Imbalanced Labels (DRIL) algorithm is proposed to decouple the highly coupled diagnostic results and rebalance the datasets, and then the balanced datasets is used to train the classifier. This paper evaluates the proposed DDN using Chinese Obstetric EMR (COEMR) datasets, and verifies the effectiveness and universality of the model on two benchmark multi-label text classification datasets: Arxiv Academic Papers Datasets (AAPD) and Reuters Corpus1 (RCV1). Demonstrating the effectiveness of the proposed methods is an imbalanced obstetric EMRs. The accuracy of DDN model on COEMR, AAPD and RCV1 datasets is 84.17, 86.35 and 93.87% respectively, which is higher than the current optimal experimental results.
Citation: Kunli Zhang, Shuai Zhang, Yu Song, Linkun Cai, Bin Hu. Double decoupled network for imbalanced obstetric intelligent diagnosis[J]. Mathematical Biosciences and Engineering, 2022, 19(10): 10006-10021. doi: 10.3934/mbe.2022467
[1] | Salman Safdar, Calistus N. Ngonghala, Abba B. Gumel . Mathematical assessment of the role of waning and boosting immunity against the BA.1 Omicron variant in the United States. Mathematical Biosciences and Engineering, 2023, 20(1): 179-212. doi: 10.3934/mbe.2023009 |
[2] | Yukihiko Nakata, Ryosuke Omori . The change of susceptibility following infection can induce failure to predict outbreak potential by R0. Mathematical Biosciences and Engineering, 2019, 16(2): 813-830. doi: 10.3934/mbe.2019038 |
[3] | Xi-Chao Duan, Xue-Zhi Li, Maia Martcheva . Dynamics of an age-structured heroin transmission model with vaccination and treatment. Mathematical Biosciences and Engineering, 2019, 16(1): 397-420. doi: 10.3934/mbe.2019019 |
[4] | Kento Okuwa, Hisashi Inaba, Toshikazu Kuniya . Mathematical analysis for an age-structured SIRS epidemic model. Mathematical Biosciences and Engineering, 2019, 16(5): 6071-6102. doi: 10.3934/mbe.2019304 |
[5] | Mostafa Adimy, Abdennasser Chekroun, Claudia Pio Ferreira . Global dynamics of a differential-difference system: a case of Kermack-McKendrick SIR model with age-structured protection phase. Mathematical Biosciences and Engineering, 2020, 17(2): 1329-1354. doi: 10.3934/mbe.2020067 |
[6] | Desmond Z. Lai, Julia R. Gog . Waning immunity can drive repeated waves of infections. Mathematical Biosciences and Engineering, 2024, 21(2): 1979-2003. doi: 10.3934/mbe.2024088 |
[7] | Lili Liu, Xi Wang, Yazhi Li . Mathematical analysis and optimal control of an epidemic model with vaccination and different infectivity. Mathematical Biosciences and Engineering, 2023, 20(12): 20914-20938. doi: 10.3934/mbe.2023925 |
[8] | Sarafa A. Iyaniwura, Rabiu Musa, Jude D. Kong . A generalized distributed delay model of COVID-19: An endemic model with immunity waning. Mathematical Biosciences and Engineering, 2023, 20(3): 5379-5412. doi: 10.3934/mbe.2023249 |
[9] | Simone De Reggi, Francesca Scarabel, Rossana Vermiglio . Approximating reproduction numbers: a general numerical method for age-structured models. Mathematical Biosciences and Engineering, 2024, 21(4): 5360-5393. doi: 10.3934/mbe.2024236 |
[10] | Junyuan Yang, Rui Xu, Xiaofeng Luo . Dynamical analysis of an age-structured multi-group SIVS epidemic model. Mathematical Biosciences and Engineering, 2019, 16(2): 636-666. doi: 10.3934/mbe.2019031 |
Electronic Medical Record (EMR) is the data basis of intelligent diagnosis. The diagnosis results of an EMR are multi-disease, including normal diagnosis, pathological diagnosis and complications, so intelligent diagnosis can be treated as multi-label classification problem. The distribution of diagnostic results in EMRs is imbalanced. And the diagnostic results in one EMR have a high coupling degree. The traditional rebalancing methods does not function effectively on highly coupled imbalanced datasets. This paper proposes Double Decoupled Network (DDN) based intelligent diagnosis model, which decouples representation learning and classifier learning. In the representation learning stage, Convolutional Neural Networks (CNN) is used to learn the original features of the data. In the classifier learning stage, a Decoupled and Rebalancing highly Imbalanced Labels (DRIL) algorithm is proposed to decouple the highly coupled diagnostic results and rebalance the datasets, and then the balanced datasets is used to train the classifier. This paper evaluates the proposed DDN using Chinese Obstetric EMR (COEMR) datasets, and verifies the effectiveness and universality of the model on two benchmark multi-label text classification datasets: Arxiv Academic Papers Datasets (AAPD) and Reuters Corpus1 (RCV1). Demonstrating the effectiveness of the proposed methods is an imbalanced obstetric EMRs. The accuracy of DDN model on COEMR, AAPD and RCV1 datasets is 84.17, 86.35 and 93.87% respectively, which is higher than the current optimal experimental results.
The dynamics of immune status among host individuals plays a crucial role in the spread of infectious diseases. Individuals which have just recovered from disease obtain any immunity, however, its level of effectiveness is not necessarily time-constant but can vary as time evolves. An individual's immunity may decay as time goes by, while it could be regained by boosting, which means the immunity enhancement by continued or intermittent exposure to infectious agent.
The boosting and waning dynamics of immune status and its epidemiological consequences have been studied by many authors. Historically speaking, it should be firstly noted that Kermack and McKendrick [2] had already developed a class-age structured epidemic model allowing reinfection of recovered individuals from the 1930s. Reinfection of recovered individuals can be seen as a consequence of waning immunity, and it can cause a backward bifurcation of endemic steady state ([3,4]). If the loss of immunity of recovered individuals implies the direct reversion from the recovered class to the susceptible class, the basic system is formulated as the well-known SIRS model, which has been long studied in mathematical epidemiology. For example, Bhattacharya and Adler [5] developed a recovery-age dependent SIRS model to show that loss of immunity could produce oscillations. On the other hand, Nakata, et al. [6] considered an infection-age structured SIRS model to examine global stability conditions for an endemic equilibrium. Okuwa, et al. [7] studied a chronological age-structured SIRS model to establish its endemic threshold results.
It is more challenging to consider synergistic consequences of boosting and waning of immunity. Heffernan and Keeling [8] discussed how vaccination can have a range of unexpected consequences as it reduces the natural boosting of immunity as well as reducing the number of naive susceptibles. Kababi{\'r}e, et al. [9] examined a model for a vector-borne disease with immunity decay and immunity boosting of human hosts. Leung, et al. [10] studied how immune boosting and cross immunity can influence the timing and severity of epidemics. It was also shown that the epidemiological patterns of an infectious disease may change considerably when the duration of vaccine-acquired immunity differs from that of infection-acquired immunity [11]. Dafilis, et al. [12] showed that allows for boosting of immunity may naturally give rise to undamped oscillatory behavior for biologically realistic parameter choices. Diekmann, et al. [13] developed a mathematical model to describe the distribution of immune status shaped by continuous waning and occasional boosting. Epidemiological consequences of waning of vaccine-derived immunity are also discussed in [14], which topic should be paid more attention in the context of COVID-19.
In a series of papers from the 1980s, Aron ([15,16,17]) developed mathematical models for malaria to consider the effect of immunity boosting by reinfection, and discussed that the boosting mechanism may explain the age-specific prevalence of acute malaria and the functional relationship between the rate of reversion from the immune class to the susceptible class and the force of infection at the endemic steady state, although those pioneering works have been so far almost neglected. In those models, the host population are divided into three classes: susceptibles (uninfected individuals); infecteds (the individuals with severe and symptomatic infection); immunes (the individuals with mild, subclinical, or asymptomatic infection). The recovered individuals are assumed to be partially susceptible and infective, and their immune status can be boosted by reinfection. Since subclinical individuals are assumed to become completely susceptible again with a recovery-age dependent reversion rate, so the Aron model can be seen as an extension of the well-known SIRS epidemic model.
In [4], Inaba showed that the Aron model can be more generally formulated as an age-structured SIRS model, where asymptomatic infecteds are distributed by the recovery-age (or immunity clock), that is, time elapsed since the moment of recovery from the symptomatic infectious class. In Inaba's formulation, the boosting effect is expressed by a boundary condition such that the immunity clock is reset to zero by boosting (reinfection). Under this formulation, the functional relationship between the reversion rate and the force of infection is then naturally induced, and Aron's result is obtained as a special case in which the probability density function for the loss of immunity is given by a delta function. If the immunity level is waning monotonically as time evolves, it is inversely proportional to recovery-age, that is, the reset of the immunity clock means the boosting of immunity.
By using the immunity clock mechanism, we could consider the dynamics of continuous change of individual immune status in the spread of infectious diseases, which is more advantageous and natural than the traditional compartment model formulation. For another kind of structured boosting models, the reader may refer to [18], [19] and [20], where immune individuals are structured by a general immunity level parameter. Although the immunity clock mechanism may be more restrictive than the general immunity level parameter dynamics, the clock mechanism is theoretically much more tractable.
For the Aron–Inaba formulation, a crucial assumption is that newly boosted individuals have the same immunity level as individuals who have just recovered from symptomatic infection, but it would be a restrictive assumption in order to consider the incomplete boosting. Therefore, in this paper, we extend the Aron–Inaba model so that the effect of boosting is that the immunity clock can be reset to any time less than the recovery-age at which reinfection occurs. If the immunity level is monotone decreasing with respect to the recovery-age, our assumption implies that newly boosted (reinfected) individuals could get, with a given probability, any level of immunity less than the maximum level that is gained by recovery from symptomatic infection status.
In the following, we first give the well-posedness result of our basic system. Next, we investigate the initial invasion condition which is formulated by the local stability of disease-free steady state. Thirdly we consider the existence of endemic steady states. Finally, we provide a bifurcation analysis of endemic steady states. Based on Lyapunov–Schmidt type arguments, we determine the direction of bifurcation that endemic steady states bifurcate from the disease-free steady state when the basic reproduction number passes through the unity. We give a necessary and sufficient condition for backward bifurcation to occur.
We consider a closed population with the constant total size N divided into three parts: susceptibles, clinical (symptomatic) infectives and subclinical (asymptomatic) infectives. The subclinical class is structured by variable τ, called the recovery-age, which is the time elapsed since the recovery from clinical infectious class. Let S(t) and I(t) be the size of susceptible and clinical infectives at time t, respectively, and let r(t,τ) be the age density function of subclinical infectious class. By extending the Aron–Inaba model [4], we formulate the following age-structured epidemic model:
dS(t)dt=μN−μS(t)−λ(t)S(t)+∫∞0γ(τ)r(t,τ)dτ,dI(t)dt=λ(t)S(t)−(μ+q)I(t),∂r(t,τ)∂t+∂r(t,τ)∂τ=−(λ(t)+μ+γ(τ))r(t,τ)+λ(t)∫∞τp(τ,σ)r(t,σ)dσ,r(t,0)=qI(t)+λ(t)∫∞0p(0,σ)r(t,σ)dσ, | (2.1) |
where μ>0 is the natural death rate, q>0 is the recovery rate from clinical infection, γ(τ) is the reversion rate at recovery-age τ. The force of infection λ is given by
λ(t)=1N(β1I(t)+∫∞0β2(τ)r(t,τ)dτ), | (2.2) |
where β1 and β2(τ) are transmission coefficients reflecting the infectivity for clinical individual and subclinical individual at recovery-age τ, respectively. For simplicity, we assume that the same force of infection λ is applied to both susceptibles and subclinical individuals.
As a technical but biologically reasonable assumption, we suppose that γ, β2∈L∞(0,∞) and denote their upper bound as γ∞ and β∞2, respectively. Moreover, we also define the survival function Γ induced from the reversion rate γ as follows:
Γ(τ):=exp(−∫τ0γ(z)dz) | (2.3) |
The parameter p(τ,σ) is a key to formulate our novel idea, which means the proportion that subclinical individuals with recovery-age σ reset their recovery-age to τ<σ when their immunity is boosted by reinfection. Then we assume the following condition:
p(0,σ)+∫σ0p(τ,σ)dτ=1,∀σ>0, | (2.4) |
so that the number of subclinical population conserves through the boosting, that is,
∫∞0r(t,σ)dσ=∫∞0∫∞τp(τ,σ)r(t,σ)dσdτ+∫∞0p(0,σ)r(t,σ)dσ, |
holds. Furthermore, we assume p∞:=supτ<σp(τ,σ)<∞. For instance, following types of function satisfy the condition (2.4):
1. p(τ,σ)=e−(σ−τ).
2.
p(τ,σ)={√2πs2exp(−(σ−τ)22s2),τ>0,1−√2πs2∫σ0exp(−η22s2)dη,τ=0, | (2.5) |
where s>0 is a given constant.
3. p(τ,σ)=11+σ.
4. p(0,σ)=0 and p(⋅,σ) is an arbitrary probability density function whose support is [0,σ].
Note that the original Aron model corresponds to the special case that p(0,σ)=1 and p(τ,σ)=0 for all τ>0 ([4], chapter 8). As is mentioned above, if the immunity level is monotone decreasing with respect to recovery-age, our assumption implies that, by through the reset of recovery-age, the immunity level of reinfected individuals could be boosted to any level of immunity less than the highest immunity level gained by recovery from the clinical status. Finally it is easily checked that
N=S(t)+I(t)+∫∞0r(t,τ)dτ, | (2.6) |
is constant for all t≥0, since we assume there is no disease-induced death rate.
As a preliminary study for our basic system (2.1), we here consider the linearized system at the diesease-free steady state (S,I,r)=(N,0,0). Let (Y,z(⋅)) be the perturbation of the infected population densities (I,r) from the equilibrium state (0,0). Then the linearized system at the disease-free steady state is described as follows:
dY(t)dt=λ(t)N−(μ+q)Y(t),∂z(t,τ)∂t+∂z(t,τ)∂τ=−(μ+γ(τ))z(t,τ),z(t,0)=qY(t),λ(t)=1N(β1Y(t)+∫∞0β2(τ)z(t,τ)dτ), | (2.7) |
which is already introduced in [4], because the boosting effect is the second order effect, so the parameter p does not appear in (2.7), and the effect of immunity boosting can be neglected as long as we consider the initial invasion phase.
Let us introduce the renewal equation for the density of newly infecteds denoted by v(t):=Nλ(t). The linearized equation for clinically infected population density in (2.7) implies
Y(t)=e−(μ+q)tY(0)+∫t0e−(μ+q)(t−s)λ(s)Nds. | (2.8) |
Integrating the second equation in (2.7) along the characteristic line, we obtain
z(t,τ)={qY(t−τ)e−μτΓ(τ),t−τ>0,z(0,τ−t)Γ(τ)Γ(τ−t)e−μτ,t−τ<0. | (2.9) |
Then we obtain the following integral equation:
v(t)=β1Y(t)+∫∞0β2(τ)z(t,τ)dτ=g(t)+β1∫t0e−(μ+q)(t−s)v(s)ds+∫t0β2(τ)qe−μτΓ(τ)∫t−τ0e−(μ+q)(t−τ−s)v(s)dsdτ, | (2.10) |
where the initial data is given by
g(t):=β1e−(μ+q)tY(0)+∫t0β2(τ)qe−μτΓ(τ)e−(μ+q)(t−τ)Y(0)dτ+∫∞tqβ2(τ)Γ(τ)Γ(τ−t)e−μtz(0,τ−t)dτ. |
Changing the order of integrals, we have the following renewal equation:
v(t)=g(t)+∫t0Φ(s)v(t−s)ds, | (2.11) |
where the net reproduction kernel Φ is defined as
Φ(s):=β1e−(μ+q)s+∫s0β2(τ)qe−μτΓ(τ)e−(μ+q)(s−τ)dτ. | (2.12) |
Then the basic reproduction number R0 is given by
R0=∫∞0Φ(s)ds=β1μ+q+qμ+q∫∞0β2(τ)e−μτΓ(τ)dτ, | (2.13) |
which is the expected number of secondary cases produced by an infected individual during its entire period of infectiousness in a completely susceptible population.
Note that the basic reproduction number is traditionally defined as the spectral radius of the next generation operator ([4,21,22]). Since the density of newly infecteds in the initial invasion phase is described by the renewal integral equation as (2.11), its next generation operator is given by
K:=∫∞0Φ(s)ds, | (2.14) |
which is acting on the state space of infected individuals. In our case, K is a scalar, so K itself gives R0.
From the epidemiological threshold principle, we can expect that if R0>1, the disease can invade into the host population, while it cannot if R0<1. We will give a mathematical justification for this invasion principle for our system in section 4.
Here we prove the existence, uniqueness and positivity of solution semiflow of the system (2.1). If I and r(⋅) are once determined, S is given by the relation (2.6). So it is sufficient to consider the (I,r)-system as follows:
dI(t)dt=λ(t)(N−I(t)−∫∞0r(t,τ)dτ)−(μ+q)I(t),∂r(t,τ)∂t+∂r(t,τ)∂τ=−(λ(t)+μ+γ(τ))r(t,τ)+λ(t)∫∞τp(τ,σ)r(t,σ)dσ,r(t,0)=qI(t)+λ(t)∫∞0p(0,σ)r(t,σ)dσ, | (3.1) |
where the state space of (I,r) is given by {(I,r)∈R+×L1+(R+):I+‖r‖L1≤N}.
In order to formulate the basic system (3.1) as a semilinear abstract Cauchy problem ([23,24,25]). Let us define the extended state space Z as Z=R×R×L1(0,∞) and its closed subspace Z0={0}×R×L1(0,∞), where the first element corresponds to the boundary value of r, and the second and third parts give the states of I and r.
Let A be a differential operator on X:=R×L1(R+) such as
(Aψ)(τ)=(−μψ1−(μ+γ(τ))ψ2−ddτψ2(τ)) for ψ=(ψ1ψ2(⋅))∈D(A):=R×W1,1(R), | (3.2) |
and define the linear operator A acting on Z such that
A(0ψ)=(−ψ2(0)Aψ) for (0ψ)∈D(A):={0}×D(A). | (3.3) |
Next define the nonlinear operator F:Z0→Z as
F(0ψ)=(qψ1+λ[ψ]∫∞0p(0,η)ψ2(η)dη(λ[ψ](N−ψ1−‖ψ2‖L1)−qψ1−λ[ψ]ψ2(τ)+λ[ψ]∫∞τp(τ,η)ψ2(η)dη)), | (3.4) |
where λ[⋅]:R+×L1(R+)→R+ is a functional defined as
λ[ψ]=1N(β1ψ1+∫∞0β2(τ)ψ2(τ)dτ). | (3.5) |
Then it is easy to see that F is a Lipschitz continuous nonlinear perturbation, that is, there is a number L>0 such that ‖Fu−Fv‖Z≤L‖u−v‖Z for all u,v∈Z0.
Now formally the system (3.1) can be written as a semilinear abstract Cauchy problem in Z:
du(t)dt=Au(t)+F(u(t)),u(0)=(0u0)∈Z0. | (3.6) |
From biological meaning, we are interested in the solution semiflow in C0⊂Z0, where
C0:={(0(ψ1ψ2(⋅)))∈Z0+:|ψ1|+‖ψ2‖L1≤N}. | (3.7) |
Then A is not densely defined in Z, but its domain is dense in Z0, that is, ¯D(A)=Z0. Moreover, we can observe that A is a resolvent positive, Hille-Yosida type operator as follows:
Lemma 1. 1. {ζ∈C∣ℜζ>−μ}⊂ρ(A).
2. For ζ>−μ, the Hille-Yosida estimation holds as
‖(ζ−A)−1‖B(Z)≤1ζ+μ | (3.8) |
3. For ζ>0, it holds that ζ(ζ−A)−1(C)⊂C0.
Proof. Let ℜζ>−μ. Consider the following resolvent equation
(ζ−A)v=u,v=(0(v1v2(⋅)))∈D(A),u=(u3(u1u2(⋅)))∈Z. |
Then v2(0)=u3, ζv1+μv1=u1 and ζv2(τ)+(μ+γ(τ))v2(τ)+v′2(τ)=u2(τ) hold. These imply
(ζ−A)−1u=(0(u1ζ+μu3e−(ζ+μ)τΓ(τ)+∫τ0u2(σ)e−(ζ+μ)(τ−σ)Γ(τ)Γ(σ)dσ)). |
Hence for ζ>−μ and u∈Z
‖(ζ−A)−1u‖Z≤|u1|ζ+μ+|u3|ζ+μ+∫∞0∫τ0|u2(σ)|e−(ζ+μ)(τ−σ)dσ=‖u‖Zζ+μ. |
From the above, it is obvious that ζ(ζ−A)−1(C)⊂C0 for given ζ>0.
Subsequently we can observe the quasi-positivity of the nonlinear perturbation term F:
Lemma 2. For sufficiently small ϵ>0, (I+ϵF)(C0)⊂C.
Proof. Let ψ=(0(ψ1ψ2(⋅)))∈C0, then we have
(I+ϵF)(0ψ)=(ϵ(qψ1+λ[ψ]∫∞0p(0,η)ψ2(η)dη)((1−ϵq)ψ1+ϵλ[ψ](N−ψ1−‖ψ2‖L1)(1−ϵλ[ψ])ψ2(τ)+ϵλ[ψ]∫∞τp(τ,η)ψ2(η)dη)). |
Since the estimation
λ[ψ]≤max{β1,β∞2}N‖ψ‖≤max{β1,β∞2}, |
holds for ψ∈C0, if we take ϵ>0 such as
0<ϵ<min{1max{β1,β∞2},1q}=:ϵ0, |
then (I+ϵF)(C0)⊂Z+. Suppose 0<ϵ<ϵ0. Due to the positivity of each element, we obtain
‖(I+ϵF)(0ψ)‖Z=ψ1+‖ψ2‖+ϵλ[ψ](N−ψ1−‖ψ2‖L1)=(ψ1+‖ψ2‖)(1−ϵλ[ψ])+ϵλ[ψ]N, |
where the condition (2.4) is used. Moreover, since 1−ϵλ[ψ]>1−λ[ψ]max{β1,β∞2}≥0, the following estimation holds:
‖(I+ϵF)(0ψ)‖Z≤N(1−ϵλ[ψ])+ϵλ[ψ]N=N, |
which implies (I+ϵF)(C0)⊂C.
In order to construct a semiflow in C0, we rewrite (3.6) as follows:
ddtu(t)=(A−1ϵ)u(t)+1ϵ(I+ϵF)u(t),u(0)∈C0, | (3.9) |
where ϵ is a small positive number such that I+ϵF maps C0 into C. Hence it is expected that the solution semiflow of (3.9) could be obtained by a positive iteration for the variation of constants formula, but the operator A has a non-dense domain in Z, A−1/ϵ is not an infinitesimal generator of a strongly continuous semigroup. Thus we seek a solution in a weak sense.
For a given T>0, u:[0,T)→C0 is an integral solution of (3.9) if u is continuous on [0,T) and it satisfies
∫t0u(s)ds∈D(A),∀t∈[0,T), | (3.10) |
and
u(t)=u(0)+(A−1ϵ)∫t0u(s)ds+1ϵ∫t0(I+ϵF)u(s)ds,∀t∈[0,T). | (3.11) |
Note that (3.10) implies that u(t)∈Z0=¯D(A) and the integral solution becomes the classical solution if and only if u is differentiable on (0,T) [26].
Let A0 be the part of A in Z0, that is,
A0=A on D(A0)={(0ψ)∈D(A):A(0ψ)∈Z0}. | (3.12) |
Then the following holds:
Lemma 3. ¯D(A0)=Z0 holds and A0 generates a strongly continuous semigroup {T0(t)}t≥0 on Z0 and T0(t)C0⊂C0 for all t>0.
Proof. It follows from definition that (0ψ)∈D(A0) if and only if ψ∈D(A) with ψ2(0)=0. Then we have ¯D(A0)=Z0. Define a differential operator A0 such that
(A0ψ)(τ)=(−μψ1−(μ+γ(τ))ψ2−ddτψ2(τ)), | (3.13) |
where
ψ=(ψ1ψ2)∈D(A0):={ψ∈D(A):ψ2(0)=0}. |
Then A0 is a generator of C0-semigroup T0(t), t≥0 on R×L1(R+) such that
T0(t)ψ=(ψ1e−μtU(t)ψ2), | (3.14) |
where U(t), t≥0 denotes the translation semigroup on L1 defined by
(U(t)ϕ)(τ):={0,t−τ>0ϕ(τ−t)e−μtΓ(τ)Γ(τ−t),τ−t>0,ϕ∈L1(R+), | (3.15) |
and we have D(A0)={0}×D(A0). Thus A0 is densely defined on Z0 and it generates a C0-semigroup on Z0 such that
T0(t)(0ψ)=(0T0(t)ψ), | (3.16) |
Since the semiflow T0(t), t>0 makes a closed subset {ψ=(ψ1,ψ2)∈R+×L1+(R+):|ψ1|+‖ψ2‖L1≤N} positively invariant, we have for t>0, (T0(t))(C0)⊂C0.
Proposition 1. For a given T>0 and u(0)∈C0, a positive function u∈C(0,T;C0) is an integral solution for (3.9) if and only if it is the positive continuous solution of the extended variation of constants formula on C0 as
u(t)=e−1ϵtT0(t)u(0)+limζ→∞1ϵ∫t0e−1ϵ(t−s)T0(t−s)ζ(ζ−(A−1ϵ))−1(I+ϵF)u(s)ds. | (3.17) |
Proof. First we assume that u∈C(0,T;C0) is an integral solution of (3.9). Applying the resolvent operator ζ(ζ−(A−1ϵ))−1 with ζ>0 to the both sides of (3.9), we obtain
uζ(t)=uζ(0)+ζ(ζ−(A−1ϵ))−1(A−1ϵ)∫t0u(s)ds+ζ(ζ−(A−1ϵ))−11ϵ∫t0(I+ϵF)u(s)ds, | (3.18) |
where uζ(t):=ζ(ζ−(A−1ϵ))−1u(t). Then it is easily seen that uζ∈D(A0)∩C0. Since ζ(ζ−(A−1ϵ))−1(A−1ϵ) is bounded, it follows that
ζ(ζ−(A−1ϵ))−1(A−1ϵ)∫t0u(s)ds=∫t0ζ(ζ−(A−1ϵ))−1(A−1ϵ)u(s)ds=∫t0(A0−1ϵ)ζ(ζ−(A−1ϵ))−1u(s)ds=∫t0(A0−1ϵ)uζ(s)ds. |
On the other hand, due to the continuity of (I+ϵF)u(t), integral and the resolvent operator can be interchanged in the third part of the right hand side in (3.18). Therefore we arrive at the following equation on D(A0)∩C0:
uζ(t)=uζ(0)+∫t0(A0−1ϵ)uζ(s)ds+1ϵ∫t0ζ(ζ−(A−1ϵ))−1(I+ϵF)u(s)ds. | (3.19) |
Then it follows that uζ is a classical solution of the Cauchy problem on Z0 as
ddtuζ(t)=(A0−1ϵ)uζ(s)+1ϵζ(ζ−(A−1ϵ))−1(I+ϵF)u(t),uζ(0)=ζ(ζ−(A−1ϵ))−1u(0). | (3.20) |
Applying the standard variation of constants formula to (3.19), we have
u(t)=e−1ϵtT0(t)u(0)+1ϵ∫t0e−1ϵ(t−s)T0(t−s)ζ(ζ−(A−1ϵ))−1(I+ϵF)u(s)ds. | (3.21) |
Therefore, it follows from limζ→∞uζ=u that we have (3.17). Conversely, we assume that u(t) is the positive continuous solution of (3.17). Consider a linear inhomogeneous Cauchy problem on Z as:
ddtv(t)=(A−1ϵ)v(t)+1ϵ(I+ϵF)u(t),v(0)=u(0). | (3.22) |
Since the generator A−1ϵ satisfies the Hille–Yosida estimate, it follows from Da Prato–Sinestrari theorem [27] that (3.22) has a unique integral solution v(t)∈Z such that
v(t)=u(0)+(A−1ϵ)∫t0v(s)ds+1ϵ∫t0(I+ϵF)v(s)ds,t∈[0,T). | (3.23) |
Applying the same argument as above, we obtain the extended variation of constants formula again as:
v(t)=e−1ϵtT0(t)u(0)+limζ→∞1ϵ∫t0e−1ϵ(t−s)T0(t−s)ζ(ζ−(A−1ϵ))−1(I+ϵF)u(s)ds, | (3.24) |
which shows that u(t)=v(t) for all t∈[0,T). Then the positive continuous solution of (3.17) is the integral solution of the basic system (3.9).
From the above arguments, we obtain the well-posedness result for the basic system (3.6) as follows:
Proposition 2. The basic model (3.6) has a unique, global, positive integral solution.
Proof. From the above observation, it is sufficient to show that the extended variation of constants formula (3.17) has a unique continuous solution for some T>0. Define an operator H on C(0,T;C0) by
(Hu)(t):=e−1ϵtT0(t)u(0)+limζ→∞1ϵ∫t0e−1ϵ(t−s)T0(t−s)ζ(ζ−(A−1ϵ))−1(I+ϵF)u(s)ds. | (3.25) |
Since the right hand side of (3.25) is a linear convex combination of elements in C0, so H maps C0 into itself. Therefore, using the local Lipschitz continuity of the bounded perturbation F and the contraction mapping principle to H, we can find a small T>0 such that a positive local solution for the extended variation of constants formula (3.17) uniquely exists. Moreover, it follows from the fact ‖T0(t)‖≤1 and ‖ζ(ζ−(A−1ϵ))−1‖≤1 that
‖(Hu)(t)‖Z≤‖u(0)‖Z+1ϵ‖I+ϵF‖∫t0‖u(s)‖Zds. |
Hence we know that the fixed point of H grows at most exponentially as time evolves, so the local solution can be extended to a global solution.
From the above argument, we know that for each u(0)∈C0, (3.9) has a unique integral solution u:R+→C0, then T(t)u(0)=u(t) defines the nonlinear semigroup T(t), t≥0 on C0. Here we consider the stability of steady states of T(t), t≥0, which are time-independent solutions of (3.11)
Here we examine the initial invasion phase of epidemic, that is, we consider the local stability of the disease-free equilibrium u=0 of T(t), t≥0. Since F is continuously F\'rechet differentiable at a steady state u=u∗, the nonlinear semigroup T(t), t≥0 is also Fréchet differentiable at u=u∗ and its derivative at u=u∗ is the strongly continuous linear semigroup S(t), t≥0 generated by the part of B:=A+F′[u∗] in Z0. Let B0 be the part of B in Z0. Then the growth bound ω0(B) and the essential growth bound ω1(B) are defined as
ω0(B0)=limt→∞t−1log(‖S(t)‖),ω1(B0)=limt→∞t−1log(α[S(t)]), | (4.1) |
if these exist, where α[A] is the measure of noncompactness of an operator A. Let s(B0):=supζ∈σ(B0)ℜζ be the spectral bound of B0. Then the following holds ([28,29]):
ω0(B0)=max{ω1(B0),s(B0)}. | (4.2) |
If ω0(B0)<0, we can conclude that the steady state u=u∗ is locally asymptotically stable, in the sense that for ω∈(ω0(B0),0) there exists δ>0 such that ‖T(t)u−u∗‖Z≤eωt‖u−u∗‖Z for all u∈C0 with ‖u−u∗‖Z≤δ. On the other hand, u∗ is unstable if at least one eigenvalue of B0 has strictly positive real part (Theorem 4.2, Corollary 4.3 in [25]).
Here let us consider the stability of the disease-free steady state for (3.6). The linearized system of (3.6) at u=0 is given as follows:
ddtu(t)=(A+DF[0])u(t), | (4.3) |
where the Fréchet derivative of F at 0∈Z is denoted by DF[0] and is calculated as
DF[0](0ψ)=(qψ1(λ[ψ]N−qψ10)). | (4.4) |
Since the range of DF[0] is finite, it is a compact operator. In the following we investigate the spectrum of the part B0 of the linearized operator B=A+DF[0].
Lemma 4. Let Pσ(B0) be the point spectrum of B0 and let Σ:={z∈C:ℜz>−μ}. Then it follows that σ(B0)∩Σ=Pσ(B0)∩Σ.
Proof. Consider the resolvent equation
(ζ−B0)v=u,v=(0(v1v2(⋅)))∈D(A),u=(0(u1u2(⋅)))∈Z0. |
Then we have
v2(0)=qv1,(ζ+μ+q)v1=u1+λ[v]N,ddτv2(τ)=−(ζ+μ+γ(τ))v2(τ)+u2(τ). |
Solving the above ODE, we have
v2(τ)=qv1e−(ζ+μ)τΓ(τ)+∫τ0u2(σ)e−(ζ+μ)(τ−σ)Γ(τ)Γ(σ)dσ. | (4.5) |
Moreover, formally we can calculate v1 as
v1=1(ζ+μ+q)(1−Δ(ζ))[u1+∫∞0β2(τ)∫τ0u2(σ)e−(ζ+μ)(τ−σ)Γ(τ)Γ(σ)dσdτ], | (4.6) |
where
Δ(ζ)=1ζ+μ+q[β1+q∫∞0β2(τ)e−(ζ+μ)τΓ(τ)dτ]. | (4.7) |
Inserting the expression of (4.6) into (4.5), v2 can be expressed by u1, u2, so we can formally calculate (ζ−B0)−1u. It is clear that (ζ−B0)−1u∈D(A) if u∈Z0 and ℜζ>−μ and Δ(ζ)≠1, so {ζ∈C:Δ(ζ)≠1,ℜζ>−μ}⊂ρ(B0). Then we have Ω:={ζ∈C:Δ(ζ)=1,ℜζ>−μ}⊃σ(B0)∩Σ. On the other hand, if ζ∈Ω, we can calculate an eigenfunction for B0 associated with ζ as (0,v1,qv1e−(μ+ζ)τΓ(τ)) where v1 is any nonzero constant, so we have ζ∈Pσ(B0). Thus we have Pσ(B0)∩Σ⊃Ω⊃σ(B0)∩Σ, but trivially it follows that Pσ(B0)∩Σ⊂σ(B0)∩Σ, hence we obtain σ(B0)∩Σ=Ω=Pσ(B0)∩Σ.
Lemma 5. If Δ(−μ)≥1, Ω={ζ∈C:Δ(ζ)=1,ℜζ>−μ} has a dominant real normal eigenvalue ζ0 and it follows that sign(ζ0)=sign(Δ(0)−1).
Proof. Since ζ↦Δ(ζ) is an analytic function in Σ, each ζ∈Ω is a normal eigenvalue. Δ(ζ) is monotone decreasing from Δ(−μ) to zero as ζ moves from −μ to ∞. Then there is a unique real root ζ0∈[−μ,∞) of Δ(ζ)=1 if Δ(−μ)≥1, and it follows that sign(ζ0)=sign(Δ(0)−1). In this case, ζ0 is the dominant eigenvalue, that is, ℜζ<ζ0 for any ζ∈Ω∖{ζ0}. In fact, if ζ∈Ω with ℜζ≤ζ0 and ℑζ≠0, we have Δ(ζ0)=1=|Δ(ζ)|<Δ(ℜζ) holds, so ℜζ<ζ0 is obtained because of the monotonicity of Δ on R.
Proposition 3. If R0>1, then the disease-free steady state is unstable, while if R0<1, then it is locally asymptotically stable.
Proof. It follows from (3.16) that ω1(A0)≤−μ and DF[0] is a compact perturbation, we have ω1((A+DF[0])0)=ω1(A0)≤−μ ([28,30]). Therefore, sign(ω0(B0))=sign(s(B0)). From the above Lemma 5, we know that s(B0)<0 if Δ(0)<1≤Δ(−μ). If Δ(−μ)<1, there is no eigenvalue in Σ because |Δ(ζ)|≤Δ(ℜζ)≤Δ(−μ)<1, so we have s(B0)<0 if Δ(0)<1. On the other hand, if Δ(0)>1, it follows from Lemma 5 that there exists a dominant positive eigenvalue. Thus we can conclude that the disease-free steady state is unstable if Δ(0)>1, while if Δ(0)<1, it is locally asymptotically stable. Finally we notice that Δ(0)=R0.
Let (S∗,I∗,r∗(⋅)) be a steady state. Then we have
0=μN−μS∗−λ∗S∗+∫∞0γ(τ)r∗(τ)dτ,0=λ∗S∗−(μ+q)I∗,dr∗(τ)dτ=−(μ+λ∗+γ(τ))r∗(τ)+λ∗∫∞τp(τ,σ)r∗(σ)dσ,r∗(0)=qI∗+λ∗∫∞0p(0,σ)r∗(σ)dσ, | (5.1) |
λ∗=1N(β1I∗+∫∞0β2(τ)r∗(τ)dτ). | (5.2) |
It is easy to see that S∗ and I∗ are expressed by λ∗ and r∗ as follows:
S∗=1μ+λ∗(μN+∫∞0γ(τ)r∗(τ)dτ),I∗=λ∗μ+q1μ+λ∗(μN+∫∞0γ(τ)r∗(τ)dτ). | (5.3) |
Solving the equation of r∗ in (5.1), we obtain
r∗(τ)=e−(μ+λ∗)τΓ(τ)r∗(0)+∫τ0∫∞σλ∗p(σ,ζ)r∗(ζ)dζe−(μ+λ∗)(τ−σ)Γ(τ)Γ(σ)dσ. | (5.4) |
Substituting the above expression into the boundary condition of r∗, it follows that
r∗(0)=qI∗+r∗(0)∫∞0p(0,η)L(0,η;λ∗)dη+∫∞0∫η0∫∞σλ∗p(σ,ζ)r∗(ζ)dζp(0,η)L(σ,η;λ∗)dσdη, | (5.5) |
where for σ<η, we define
L(σ,η;λ∗):=λ∗e−(μ+λ∗)(η−σ)Γ(η)Γ(σ). |
Then p(ζ,η)L(σ,η;λ∗) gives the probability density that asymptomatic individuals with recovery-age σ survive to the age of η and then returns to recovery-age ζ. Since ∫∞σp(0,η)L(σ,η;λ∗)dη<1, we have
r∗(0)=qI∗+∫∞0∫η0∫∞σλ∗p(σ,ζ)r∗(ζ)dζp(0,η)L(σ,η;λ∗)dσdη1−∫∞0p(0,η)L(0,η;λ∗)dη. | (5.6) |
Then we obtain an integral equation with respect to r∗ as
r∗(τ)=qI∗e−(μ+λ∗)τΓ(τ)1−∫∞0p(0,η)L(0,η;λ∗)dη+e−(μ+λ∗)τΓ(τ)∫∞0∫η0∫∞σλ∗p(σ,ζ)r∗(ζ)dζp(0,η)L(σ,η;λ∗)dσdη1−∫∞0p(0,η)L(0,η;λ∗)dη+∫τ0∫∞σp(σ,ζ)r∗(ζ)dζL(σ,τ;λ∗)dσ, | (5.7) |
which can be expressed as follows:
r∗(τ)=f(τ;λ∗)+∫τ0∫∞σp(σ,ζ)r∗(ζ)dζL(σ,τ;λ∗)dσ+∫∞0r∗(ζ)K1(τ,ζ;λ∗)dζ, | (5.8) |
where
g(τ;λ∗):=e−(μ+λ∗)τΓ(τ)1−∫∞0p(0,η)L(0,η;λ∗)dη,f(τ;λ∗):=qλ∗μ+qμNμ+λ∗g(τ;λ∗),K1(τ,ζ;λ∗):=g(τ;λ∗)(1μ+λ∗qλ∗μ+qγ(ζ)+∫ζ0∫∞σλ∗p(σ,ζ)p(0,η)L(σ,η;λ∗)dηdσ). |
The integral equation (5.8) can be seen as the abstract equation with respect to u∈X+:=L1+(R+):
u=fλ∗+Vλ∗u+Fλ∗u, | (5.9) |
where fλ∗(τ)=f(τ;λ∗), Vλ∗ and Fλ∗ are bounded linear operators from X+ into itself given by
(Vλ∗u)(τ):=∫τ0∫∞σp(σ,η)u(η)dηL(σ,τ;λ∗)dσ,(Fλ∗u)(τ):=∫∞0u(ζ)K1(τ,ζ;λ∗)dζ. |
Since Vλ∗ is a Volterra operator, its spectral radius is zero, so the positive operator (I−Vλ∗)−1 exists. Thus the equation (5.9) is rewritten as
u=(I−Vλ∗)−1fλ∗+(I−Vλ∗)−1Fλ∗u. | (5.10) |
Since the operator ((I−Vλ∗)−1Fλ∗) is a Volterra operator again, the solution of (5.10) is obtained by the iteration:
u=∞∑k=0((I−Vλ∗)−1Fλ∗)k(I−Vλ∗)−1fλ∗. | (5.11) |
Substituting the expression (5.11) into (5.2), we obtain the following one-dimensional equation with respect to unknown λ∗:
1=1N(β1μ+q1μ+λ∗(μN+∫∞0γ(τ)r∗(τ)dτ)+∫∞0β2(τ)r∗(τ)λ∗dτ)=β1μ+qμμ+λ∗+1N∫∞0(β1μ+qλ∗μ+λ∗γ(τ)+β2(τ))r∗(τ)λ∗dτ, | (5.12) |
where
r∗(τ)=∞∑k=0((I−Vλ∗)−1Fλ∗)k(I−Vλ∗)−1fλ∗. |
From the above characteristic equation, the following statement is obtained.
Proposition 4. If R0>1, then there exists at least one endemic steady state.
Proof. Let H(λ∗) be the right hand side of (5.12). First, we prove H(0)=R0. For this purpose the value of r∗(τ)λ∗ at λ∗=0 needs to be calculated. Since the operator F0 is a zero operator, all of the terms for k≥1 in (5.11) are zero. Then we obtain
r∗(τ)λ∗|λ∗=0=fλ∗(τ)λ∗|λ∗=0=qNμ+qe−μτΓ(τ), |
from which it follows that H(0)=R0. Next we prove limλ∗→∞H(λ∗)=0. First, observe that
H(λ∗)≤β1μ+qμμ+λ∗+1N(β1μ+qγ∞+β∞2)∫∞0r∗(τ)λ∗dτ. |
By using the equality obtained by dividing (5.9) by λ∗ the following representation follows:
∫∞0r∗(τ)λ∗dτ=∫∞0[fλ∗(τ)λ∗+Vλ∗r∗(τ)λ∗+Fλ∗r∗(τ)λ∗]dτ, | (5.13) |
where each integral of the right hand side of (5.13) can be estimated as follows:
∫∞0fλ∗(τ)λ∗dτ≤qμ+q⋅μNμ+λ∗∫∞0e−(μ+λ∗)τΓ(τ)1−∫∞0p(0,η)L(0,η;λ∗)dη≤qμ+q⋅μNμ+λ∗⋅μ+λ∗μ∫∞0e−(μ+λ∗)τdτ=qN(μ+q)(μ+λ∗), |
∫∞0Vλ∗u(τ)λ∗dτ≤∫∞0∫τ0∫∞σp(σ,η)u(η)dηe−(μ+λ∗)(τ−σ)dσdτ=∫∞0∫∞σp(σ,η)u(η)dη∫∞σe−(μ+λ∗)(τ−σ)dτdσ=1μ+λ∗∫∞0∫η0p(σ,η)dσu(η)dη≤Nμ+λ∗, |
∫∞0Fλ∗u(τ)λ∗dτ≤μ+λ∗μ∫∞0e−(μ+λ∗)τ⋅1μ+λ∗⋅qμ+qγ(τ)dτ+μ+λ∗μ∫∞0∫τ0∫∞σp(σ,τ)p(0,η)e−(μ+λ∗)(η−σ)dηdσdτ≤qγ∞μ(μ+q)(μ+λ∗)+μ+λ∗μ∫∞0e−(μ+λ∗)τ∫∞σp(σ,τ)∫∞σe−(μ+λ∗)(η−σ)dηdτdσ≤1μ(μ+λ∗)(qγ∞μ+q+1). |
Then limλ∗→∞H(λ∗)=0 holds. From the above, because of the continuity of H, it turns out that the equation 1=H(λ∗) has at least one positive solution if R0>1.
As is seen in the next section, in our basic model (2.1), R0<1 does not necessarily imply non-existence of endemic steady states. However, we can formulate a sufficient condition for the global stability of the disease-free steady state implying the non-existence of endemic steady states. Let us define a time-dependent reproduction number Re(t) at time t as
Re(t):=S(t)(μ+q)N(β1I(t)+∫∞0β2(τ)r(t,τ)dτ), | (5.14) |
which is a kind of the effective reproduction number, because it is defined in a partially immunized host population. For the use and the definition of effective reproduction number, the reader may consult [4].
From the above definition, we have
dI(t)dt=(μ+q)(Re(t)−1)I(t). | (5.15) |
Then we can expect that the disease will be eradicated if the effective reproduction number Re(t) is suppressed below the unity and it holds that
∫∞0(Re(x)−1)dx=−∞, | (5.16) |
which is a criteria for disease prevention.
Proposition 5. Suppose that (5.16) holds. Then the disease-free steady state is globally asymptotically stable, that is, there is no endemic steady state.
Proof. It follows from (5.15) that
I(t)=I(0)exp((μ+q)∫t0(Re(x)−1)dx), |
from which we know that (5.16) implies that limt→∞I(t)=0. Next let U(t):=∫∞0r(t,τ)dτ. Observe that
dU(t)dt=qI(t)−μU(t)−∫∞0γ(τ)r(t,τ)dτ≤qI(t)−μU(t). |
Then we have
U(t)≤U(0)e−μt+∫t0e−μ(t−τ)qI(τ)dτ, |
which implies that limt→∞U(t)=0. Then (S(t),I(t),r(t,τ)) converges to (N,0,0) as t→∞, so the disease-free steady state is globally asymptotically stable.
Corollary 1. If
β∞N4(μ+q)<1, | (5.17) |
then the disease-free steady state is globally asymptotically stable.
Proof. Observe that
Re(t)=S(t)(μ+q)N(β1I(t)+∫∞0β2(τ)r(t,τ)dτ)≤β∞(μ+q)NS(t)(N−S(t)), |
where we know that S(N−S)≤N2/4. If (5.17) holds, we obtain
Re(t)−1<β∞N4(μ+q)−1<0, |
which implies (5.16).
We here consider the bifurcation problem of endemic steady states from the disease-free steady state when the basic reproduction number R0 crosses the unity. Especially, the direction of bifurcation is biologically important because in the case of backward bifurcation, there exist multiple endemic steady states even when R0<1, so R0<1 is not a sufficient condition to eradicate the disease, whereas if forward bifurcation occurs at R0=1 and there exists no endemic steady state for R0<1, lowering R0 less than unity becomes the almighty policy for disease eradication. Moreover, we can conclude from the Factorization Theorem [31] that the forwardly bifurcated solution is locally asymptotically stable, while the backwardly bifurcated solution is unstable as long as |R0−1| is small enough.
First, according to [1], we apply the Lyapunov-Schmidt type arguments to the one-parameter bifurcation model. Let us consider the nonlinear operator G:Z0×R→Z with a parameter ˉβ defined as
G(u,ˉβ)(τ)=(−u2(0)+qu1+λ[u;ˉβ]∫∞0p(0,σ)u2(σ)dσλ[u;ˉβ](N−u1−‖u2‖L1)−(μ+q)u1−u′2(τ)−(λ[u;ˉβ]+μ+γ(τ))u2(τ)+λ[u;ˉβ]∫∞τp(τ,σ)u2(σ)dσ) | (6.1) |
for u=T(0,u1,u2(⋅))∈Z0, where
λ[u;ˉβ]:=ˉβN(β10u1+∫∞0β20(τ)u2(τ)dτ), | (6.2) |
where β10 and β20 are normalized transmission coefficients such that
β10μ+q+qμ+q∫∞0β20(τ)e−μτΓ(τ)dτ=1. | (6.3) |
Then it follows that ˉβ=R0 under the above assumption.
Then the system (2.1) is rewritten as the following abstract differential equation:
dudt=G(u,ˉβ). | (6.4) |
This equation has a trivial equilibrium 0∈Z0, which corresponds to the disease-free steady state. Let DuG[u,ˉβ] and D2uuG[u,ˉβ] be the first and second order derivative of G with respect to u, respectively. Similarly, let D2uˉβG[u,ˉβ] be the second order derivative of G with respect to u and ˉβ. Note that DuG[u,ˉβ] is a linear map, while D2uuG[u,ˉβ] and D2uˉβG[u,ˉβ] are bilinear maps. In order to study the direction of the bifurcation, we apply the following theorem:
Theorem 1 ([1]). Suppose that the following conditions hold:
A1 Let L=DuG[0,1] be the linearization around the zero equilibrium, evaluated at a critical value of parameter ˉβ=1, such that L is a closed operator with a simple isolated eigenvalue zero and remaining eigenvalues having negative real part. Let ˆv0 be the unique (up to a constant) positive solution of Lξ=0.
A2G(u,ˉβ)∈C2(U0×I0,Z) for some neighborhood U0×I0 of (0,1).
A3 Let Z∗ be the dual of Z and ⟨⋅,⋅⟩ be the duality pairing between Z and Z∗. Assume ˆv∗0∈Z∗ is the unique (up to a constant) positive vector satisfying ⟨Lu,ˆv∗0⟩=0, ∀u∈Z0, that is, dim(kerL∗)=1, where L∗ is the adjoint of L, and kerL∗=span{ˆv∗0}.
A4⟨D2uˉβG[0,1](ˆv0,1),ˆv∗0⟩≠0.
Then, the direction of the bifurcation is determined by the numbers
a=⟨D2uuG[0,1](ˆv0,ˆv0),ˆv∗0⟩,b=⟨D2uˉβG[0,1](ˆv0,1),ˆv∗0⟩. | (6.5) |
If b>0, the bifurcation is backward if and only if a>0 and forward if only if a<0.
For our system, the derivatives of the nonlinear mapping G are calculated as the following:
(DuG[0,ˉβ]ξ)(τ)=(−ξ2(0)+qξ1λ[ξ;ˉβ]N−(μ+q)ξ1−ξ′2(τ)−(μ+γ(τ))ξ2(τ)), |
(D2uˉβG[0,ˉβ]ξ)(τ)=(0λ[ξ;ˉβ]N0), |
(D2uuG[0,ˉβ](ξ,η))(τ)=(λ[ξ;ˉβ]∫∞0p(0,σ)η2(σ)dσ+λ[η;ˉβ]∫∞0p(0,σ)ξ2(σ)dσ−λ[ξ;ˉβ](η1+‖η2‖L1)−λ[η;ˉβ](ξ1+‖ξ2‖L1)λ[ξ;ˉβ](∫∞τp(τ,σ)η2(σ)−η2(τ))+λ[η;ˉβ](∫∞τp(τ,σ)ξ2(σ)−ξ2(τ))), |
for ξ=T(0,ξ1,ξ2(⋅))∈Z0 and η=T(0,η1,η2(⋅))∈Z0. We define the linear operator L:=DuG[0,1]. If ζ∈C is an eigenvalue of L, there exists ξ=T(0,ξ1,ξ2(⋅))∈Z0∖{0} such that
−ξ2(0)+qξ1=0,ζξ1=λ[ξ;1]N−(μ+q)ξ1,ζξ2(τ)=−ξ′2(τ)−(μ+γ(τ))ξ2(τ), | (6.6) |
Then the eigenvalue ζ is given as a root of the characteristic equation:
ζ+μ+q=β10+q∫∞0β20(τ)e−(ζ+μ)τΓ(τ)dτ. | (6.7) |
If ˉβ=1, we can use the similar discussion as in the Section 4 to show that the equation (6.7) has a single solution zero and each of the remaining solutions has negative real part. Solving Lξ=0, the positive eigenvector ˆv0∈Z corresponding to the eigenvalue 0 is obtained as
ˆv0=T(q,1,qe−μτΓ(τ)). | (6.8) |
Next we calculate the adjoint operator of L and its eigenfunctional corresponding to the dominant eigenvalue zero. For ξ=(0,ξ1,ξ2(⋅))∈Z0 and ψ=T(ψ2(0),ψ1,ψ2(⋅))∈Z∗=R×R×L∞,
⟨Lξ,ψ⟩=(−ξ2(0)+qξ1)ψ3+(λ[ξ,1]N−(μ+q)ξ1)ψ1−∫∞0(ξ′2(τ)+(μ+γ(τ))ξ2(τ))ψ2(τ)dτ=(ψ2(0)−ψ3)ξ2(0)+(qψ3−(μ+q−β10)ψ1)ξ1+∫∞0[ψ′2(τ)−(μ+γ(τ))ψ2(τ)+β20(τ)ψ1]ξ2(τ)dτ, |
hence the following representation is obtained:
L∗ψ(τ)=(ψ2(0)−ψ3qψ3−(μ+q−β10)ψ1ψ′2(τ)−(μ+γ(τ))ψ2(τ)+β20(τ)ψ1) for ψ(τ)=(ψ3ψ1ψ2(τ))∈Z∗, | (6.9) |
where Z∗ denotes the adjoint space of Z. In order to find the eigenvector of L∗ corresponding to the eigenvalue zero, we consider L∗ψ=0, that is,
ψ2(0)=ψ3,(μ+q−β10)ψ=qψ3,ψ′2(τ)=(μ+γ(τ))ψ2(τ)−β20(τ)ψ1. | (6.10) |
By our assumption (6.3), we have
μ+q−β10=q∫∞0β20(τ)e−μσΓ(σ)dσ>0. | (6.11) |
Then it follows that
ψ1=qψ3μ+q−β10. | (6.12) |
Solving the ordinary differential equation in (6.10), we have
ψ2(τ)=ψ3e−μτΓ(τ)−qψ3μ+q−β10∫τ0β20(τ)e−μ(σ−τ)Γ(σ)Γ(τ)dσ=ψ3e−μτΓ(τ)[1−qμ+q−β10∫τ0β20(σ)e−μσΓ(σ)dσ]=qψ3μ+q−β10∫∞τβ20(σ)e−μ(σ−τ)Γ(σ)Γ(τ)dσ. | (6.13) |
Therefore the unique (up to constant) eigenfunctional ˆv∗0∈Z∗ of L∗ corresponding to eigenvalue zero is obtained as
ˆv∗0=(μ+q−β10q1∫∞τβ20(σ)e−μ(σ−τ)Γ(σ)Γ(τ)dσ). | (6.14) |
From the above, it is obvious that b:=⟨D2uˉβG[0,1](ˆv0,1),ˆv∗0⟩>0, hence the bifurcation is backward if a>0, while if a<0, the bifurcation is forward.
Now calculate the value of a:=⟨D2uuG[0,1](ˆv0,ˆv0),ˆv∗0⟩. Since
D2uuG[0,1](ˆv0,ˆv0)=2λ[ˆv0;1](∫∞0p(0,η)qe−μηΓ(η)dη−(1+∫∞0qe−μτΓ(τ)dτ)∫∞τp(τ,η)qe−μηΓ(η)dη−qe−μτΓ(τ)), | (6.15) |
it follows that the value of a satisfies
a=2(μ+q)N[μ+q−β10q∫∞0p(0,η)qe−μηΓ(η)dη−(1+∫∞0qe−μτΓ(τ)dτ)+∫∞0∫∞σβ20(τ)e−μ(τ−σ)Γ(τ)Γ(σ)dτ[∫∞σp(σ,η)qe−μηΓ(η)dη−qe−μσΓ(σ)]dσ]. | (6.16) |
In order to see the meaning of the bifurcation condition by the parameter a, we adopt a more intuitive argument to find the direction of the bifurcation based on the characteristic equation (5.12). The direction of the bifurcation is determined by the sign of H′(0) when R0=1. In fact, if R0<1 but |R0−1| is small enough, H′(0)>0 implies the existence of positive solution for the equation (5.12) (Figure 1).
Then we can conclude that if H′(0)>0, the bifurcation is backward, while it is forward if H′(0)<0 at R0=1. Now we prove the equivalence of both conditions.
Proposition 6. Suppose that R0=1. Then it holds that sign(a)=sign(H′(0)).
Proof. First we calculate the derivative
ddλ∗r∗(τ)λ∗|λ∗=0 |
using the equality (5.11). For the purpose of calculating the derivative at zero, we may neglect the second or higher terms of λ∗, hence
ddλ∗r∗(τ)λ∗|λ∗=0=ddλ∗(fλ∗(τ)λ∗+V∗λfλ∗λ∗(τ)+F∗λfλ∗λ∗(τ))|λ∗=0. |
Since
ddλ∗g(τ;λ∗)|λ∗=0=−τe−μτΓ(τ)+e−μτΓ(τ)∫∞0p(0,η)e−μηΓ(η)dη |
holds, it follows that
ddλ∗fλ∗(τ)λ∗|λ∗=0=qNμ+qe−μτΓ(τ)[−τ−1μ+∫∞0p(0,η)e−μηΓ(η)dη]. |
Next,
ddλ∗Vλ∗fλ∗λ∗(τ)|λ∗=0=∫∞0∫∞σp(σ,η)f(τ;λ∗)λ∗|λ∗=0ddλ∗L(σ,τ;λ∗)|λ∗=0dηdσ=qNμ+qe−μτΓ(τ)∫τ0∫∞σp(σ,η)e−μ(η−σ)Γ(η)Γ(σ)dηdσ |
and
ddλ∗Fλ∗fλ∗λ∗(τ)|λ∗=0=∫∞0qμ+qμNμ+λ∗g(ζ;λ∗)g(τ;λ∗)(1μ+λ∗qμ+qγ(ζ)+∫ζ0∫∞σp(σ,ζ)p(0,η)L(σ,η;λ∗)dηdσ)|λ∗=0dζ=e−μτΓ(τ)Nμ(qμ+q)2∫∞0e−μηΓ(η)γ(η)dη |
are obtained. Finally we arrive at the following expression:
H′(0)=−β10μ(μ+q)+β10μ(μ+q)⋅qμ+q∫∞0γ(τ)e−μτΓ(τ)dτ+qμ+q∫∞0β20(τ)e−μτΓ(τ)[−τ−1μ+∫∞0p(0,η)e−μηΓ(η)dη+∫τ0∫∞σp(σ,η)e−μ(η−σ)Γ(η)Γ(σ)dηdσ+qμ(μ+q)∫∞0e−μηΓ(η)γ(η)dη]dτ. | (6.17) |
Now let us prove that
(μ+q)H′(0)=N2(μ+q)a. | (6.18) |
Let
(μ+q)H′(0)=H1+H2+H3+H4, |
where
H1:=−β10μ+β10μ⋅qμ+q∫∞0e−μτΓ(τ)γ(τ)dτ−qμ∫∞0β20(τ)e−μτΓ(τ)dτ,H2:=q∫∞0β20(τ)e−μτΓ(τ)dτ[∫∞0p(0,η)e−μηΓ(η)dη+qμ(μ+q)∫∞0e−μηΓ(η)γ(η)dη],H3:=q∫∞0β20(τ)e−μτΓ(τ)∫τ0∫∞σp(σ,η)e−μ(η−μ)Γ(η)Γ(σ)dηdσdτ,H4:=−q∫∞0τβ20(τ)e−μτΓ(τ)dτ. |
Note that the following holds:
∫∞0e−μτΓ(τ)γ(τ)dτ=1−μ∫∞0e−μτΓ(τ)dτ. | (6.19) |
Then each Hi can be calculated as follows:
H1=−β10μ+q(1+q∫∞0e−μτΓ(τ))−qμ∫∞0β20(τ)e−μτΓ(τ)dτ=−(1−qμ+q∫∞0β20(τ)Γ(τ)dτ)(1+q∫∞0e−μτΓ(τ))−qμ∫∞0β20(τ)e−μτΓ(τ)dτ=−(1+q∫∞0e−μτΓ(τ))−q2μ(μ+q)∫∞0β20(τ)e−μτΓ(τ)dτ+q2μ+q∫∞0β20(τ)e−μτΓ(τ)dτ∫∞0e−μτΓ(τ)dτ,H2=q∫∞0β20(τ)e−μτΓ(τ)dτ⋅[∫∞0e−μηΓ(η)(1−∫η0p(σ,η)dη)dη+qμ(μ+q)(1−μ∫∞0e−μηΓ(η)dη)]=q2μ(μ+q)∫∞0β20(τ)e−μτΓ(τ)dτ−q∫∞0β20(τ)e−μτΓ(τ)dτ∫∞0∫η0p(σ,η)dσe−μηΓ(η)dη+μqμ+q∫∞0β20(τ)e−μτΓ(τ)dτ∫∞0e−μηΓ(η)dη,H3=q∫∞0∫∞σ∫∞σβ20(τ)e−μτΓ(τ)p(σ,η)e−μ(η−σ)Γ(η)Γ(σ)dηdτdσ=∫∞0∫∞σβ20(τ)e−μ(τ−σ)Γ(τ)Γ(σ)dτ∫∞σqp(σ,η)e−μηΓ(η)dηdσ,H4=−∫∞0∫τ0dσβ20(τ)qe−μτΓ(τ)dτ=−∫∞0∫∞σβ20(τ)e−μ(τ−σ)Γ(τ)Γ(σ)dτqe−μσΓ(σ)dσ. |
Therefore we have
H′(0)=∫∞0β20(τ)e−μτΓ(τ)dτ∫∞0qe−μηΓ(η)[1−∫η0p(σ,η)dσ]dη−(1+q∫∞0e−μτΓ(τ))+∫∞0∫∞σβ20(τ)e−μ(τ−σ)Γ(τ)Γ(σ)dτ[∫∞σqp(σ,η)e−μηΓ(η)dη−qe−μσΓ(σ)]dσ=μ+q−β10q∫∞0p(0,η)qe−μηΓ(η)dη−(1+q∫∞0e−μτΓ(τ))+∫∞0∫∞σβ20(τ)e−μ(τ−σ)Γ(τ)Γ(σ)dτ[∫∞σp(σ,η)qe−μηΓ(η)dη−qe−μσΓ(σ)]dσ, | (6.20) |
which implies the equality (6.18).
In order to see under what kind of biological conditions the backward bifurcation could occur, let us define a function:
F(η):=∫∞0β20(x+η)e−μxΓ(x+η)Γ(η)dx. | (6.21) |
Then F(η) is the expected number of reproduction (reproductive value) for sub-clinically infected individuals at recovery-age η, especially F(0) gives the reproduction number of newly recovered individuals. Note that our condition (6.3) is expressed as
R0=β10μ+q+qμ+qF(0)=1. | (6.22) |
Moreover we define
G(η):=F(0)p(0,η)+∫η0F(σ)p(σ,η)dσ. | (6.23) |
Then G(η) denotes the average reproduction number of sub-clinical individuals boosted at recovery-age η.
Proposition 7. Under the assumption (6.3), backward bifurcation of endemic steady states occurs at R0=1 from the disease-free steady state if and only if
q∫∞0G(η)e−μηΓ(η)dη>q∫∞0(F(η)+1)e−μηΓ(η)dη+1. | (6.24) |
Proof. From the above arguments so far, it is sufficient to show that
H′(0)=q∫∞0G(η)e−μηΓ(η)dη−q∫∞0(F(η)+1)e−μηΓ(η)dη−1. | (6.25) |
It follows from (6.20) that
H′(0)=qJ−1−q∫∞0(1+F(η))e−μηΓ(η)dη, |
where
J:=∫∞0β20(τ)e−μτΓ(τ)dτ∫∞0e−μηΓ(η)(1−∫η0p(σ,η)dσ)dη+∫∞0∫∞σβ20(τ)e−μ(τ−σ)Γ(τ)Γ(σ)dτ∫∞σp(σ,η)e−μηΓ(η)dη=F(0)∫∞0e−μηΓ(η)(1−∫η0p(σ,η)dσ)dη+∫∞0F(σ)∫∞σp(σ,η)e−μηΓ(η)dηdσ=∫∞0F(0)p(0,η)e−μηΓ(η)dη+∫∞0∫η0F(σ)p(σ,η)dσe−μηΓ(η)dη. |
Therefore it is easy to see that (6.25) holds.
Corollary 2. Suppose that
G(η)≤F(η)+1,∀η≥0. | (6.26) |
Then the forward bifurcation of endemic steady states occurs at R0=1 from the disease-free steady state. Especially one of the following typical conditions is satisfied, then (6.26) holds:
1. F is constant,
2. supσ≥0F(σ)≤1,
3. p(0,η)=1 and F(0)≤1+F(η) for all η.
Proof. From (6.25), it is clear that (6.26) is a sufficient condition in order to satisfy H′(0)<0. If F is constant, it follows from (2.4) that G(η)−F(η)=0, hence we have (6.26). Next, from (2.4), we have G(η)≤supσ≥0F(σ), so (6.26) holds if supσ≥0F(σ)≤1. Finally, if p(0,η)=1 for all η, that is, all newly reinfected individual has recovery age zero, it follows that G(η)=F(0), so again (6.26) holds, if F(0)≤1+F(η) for all η.
Corollary 3. Suppose that ∫∞σp(σ,η)dη≤1 for all σ>0. Then, the endemic steady state bifurcates forwardly from the disease-free steady state at R0=1.
Proof. Since
R0=β10μ+q+qμ+qF(0)=1, |
we have β10=μ+q−qF(0). Since β10 is nonnegative, F(0) should satisfy
F(0)≤μ+qq. | (6.27) |
Moreover, since p(0,η)=1−∫η0p(σ,η)dσ for all η>0, we have
p(0,η)≤1 | (6.28) |
for all η>0. From (6.27) and (6.28), we have
q∫∞0F(0)p(0,η)e−μηΓ(η)dη≤(μ+q)∫∞0e−μηΓ(η)dη≤μ∫∞0e−μηdη+q∫∞0e−μηΓ(η)dη=q∫∞0e−μηΓ(η)dη+1. | (6.29) |
Moreover, we have
q∫∞0∫η0F(σ)p(σ,η)dσe−μηΓ(η)dη=q∫∞0F(σ)∫∞σp(σ,η)e−μηΓ(η)dηdσ=q∫∞0F(σ){[∫ησp(σ,ρ)dρe−μηΓ(η)]∞σ+∫∞σ[μ+γ(η)]∫ησp(σ,ρ)dρe−μηΓ(η)dη}dσ≤q∫∞0F(σ)∫∞σ[μ+γ(η)]e−μηΓ(η)dηdσ=q∫∞0F(η)e−μηΓ(η)dη. | (6.30) |
Since
G(η)=F(0)p(0,η)+∫η0F(σ)p(σ,η)dσ, |
we have from (6.29) and (6.30) that
q∫∞0G(η)e−μηΓ(η)dη≤q∫∞0[F(η)+1]e−μηΓ(η)dη+1. |
Thus, the inequality (6.24) does not hold.
Corollary 4. Suppose that F is monotone on [0,∞). Then, the endemic steady state bifurcates forwardly from the disease-free steady state at R0=1.
Proof. First, suppose that F is monotone nonincreasing on [0,∞). Since p(0,η)=1−∫η0p(σ,η)dσ for all η>0, we have
G(η)=F(0)p(0,η)+∫η0F(σ)p(σ,η)dσ=F(0)+∫η0[F(σ)−F(0)]p(σ,η)dσ≤F(0)≤μ+qq,∀η≥0. |
Hence, we have
q∫∞0G(η)e−μηΓ(η)dη≤(μ+q)∫∞0e−μηΓ(η)dη≤q∫∞0e−μηΓ(η)dη+μ∫∞0e−μηdη≤q∫∞0[F(η)+1]e−μηΓ(η)dη+1. |
Thus, the inequality (6.24) in Proposition 7 does not hold. Next, suppose that F is monotone nondecreasing on [0,∞). We then have
G(η)=F(0)p(0,η)+∫η0F(σ)p(σ,η)dσ≤F(0)p(0,η)+F(η)∫η0p(σ,η)dσ≤F(0)+F(η)≤μ+qq+F(η),∀η≥0. |
Hence, we have
q∫∞0G(η)e−μηΓ(η)dη≤(μ+q)∫∞0e−μηΓ(η)dη+q∫∞0F(η)e−μηΓ(η)dη≤q∫∞0[F(η)+1]e−μηΓ(η)dη+μ∫∞0e−μηdη≤q∫∞0[F(η)+1]e−μηΓ(η)dη+1. |
Thus, (6.24) in Proposition 7 does not hold.
Proposition 8. F(0)≤1 if and only if
∫∞0β20(τ)e−β1τΓ(τ)dτ≤1. | (6.31) |
And the condition (6.31) is satisfied if supτ≥0β20(τ)≤β10. Then if p(0,η)=1 for all η≥0, the condition (6.31) is a sufficient condition to show that the forward bifurcation of endemic steady states occurs at R0=1 from the disease-free steady state.
Proof. It follows from (6.22) that β10−μ=q(1−F(0)). Then F(0)≤1 implies β10≥μ. Hence we have
∫∞0β20(τ)e−β10τΓ(τ)dτ≤∫∞0β20(τ)e−μτΓ(τ)dτ=F(0)≤1, |
so F(0)≤1 implies (6.31). Conversely suppose that (6.31) holds. If we assume that F(0)>1, it follows that β10−μ<0, which implies
F(0)<∫∞0β20(τ)e−β1τΓ(τ)dτ≤1, |
which contradicts F(0)>1, then it must hold that F(0)≤1. Therefore, the condition (6.31) implies F(0)≤1, and it also implies the condition 3 of Corollary 2, so the bifurcation is forward. Finally note that if supτ≥0β20(τ)≤β10, then we can observe that
∫∞0β20(τ)e−β10τΓ(τ)dτ≤supτ≥0β20(τ)β10≤1. |
Then the condition (6.31) is satisfied.
From the above propositions, we know that age-dependent variability of the transmission coefficient for subclinical status is needed to produce possibility of backward bifurcation of endemic steady states at R0=1 from the disease-free steady state. Yang and Nakata [32] proved that if p(0,η)=1 for all η≥0 and the condition (6.31) is satisfied, there exists a unique endemic steady state exists if and only if R0>1. Therefore, if all newly boosted infecteds have the recovery-age zero, subcritical endemic steady state does not exist if there is no enhancement of the infectivity such that supτ≥0β20(τ)>β10 among subclinical infecteds.
In this paper, we have extended the Aron epidemic model so that the immunity clock can be reset to any time less than the recovery-age at which reinfection occurs. If we assume that the immunity level is determined by the immunity clock, our model assumption allows that newly boosted (reinfected) individuals could get any level of immunity, which is a most flexible extension for the epidemic with boosting and waning of immune status.
We have established the mathematical well-posedness of our formulation and have shown that the initial invasion condition and the endemicity can be characterized by the basic reproduction number R0. Our focus is the bifurcation analysis of endemic steady states. Based on Lyapunov–Schmidt type arguments, we have determined the direction of bifurcation that endemic steady states bifurcate from the disease-free steady state when the basic reproduction number passes through the unity. We have given a necessary and sufficient condition for backward bifurcation to occur. Although we have not yet discovered biologically reasonable, concrete numerical examples of subcritical endemic steady states for our model*, we have cleared what kind of mathematical conditions are needed to create the backward bifurcation from the disease-free steady state under the reset mechanism of the immunity clock.
*In numerical simulation, we can use a standard method such as Gauss-Laguerre quadrature to calculate numerical integrals on infinite intervals.
Mathematically speaking, even in case that the forward bifurcation occurs, it is still a difficult problem to show conditions for the uniqueness of endemic steady state. It is a future challenge to investigate the global dynamics of the epidemic model with boosting and waning of immune status. From application point of view, it would be also important to extend our recovery-age dependent model to take into account the chronological-age structure of host population. As is suggested in [4], the chronological-age dependent model is essential to understand the age-specific prevalence dynamics for the epidemic with boosting and waning of immune status.
Finally, we can again stress importance to consider the waning and boosting dynamics in the real-world application. As is strongly expected in the pandemic of COVID-19, we believe that vaccination is the most effective tool to control the pandemic. However, if the effect of vaccination is not necessarily lifelong and it will wane as time evolves, so we may well have to boost the immune status by repeating vaccination to keep the herd immunity. To make an appropriate immunization policy, we may need to know the waning and boosting dynamics of immune status under any vaccination policy. It is a future challenge to extend our model by taking into account possible vaccination policy.
Kento Okuwa and Hisashi Inaba are supported by the Japan Society for the Promotion of Science (JSPS), Grant-in-Aid for Scientific Research (C) (No.16K05266). Toshikazu Kuniya is supported by JSPS Grant-in-Aid for Early-Career Scientists (No.19K14594).
There is no conflict of interest.
[1] |
Y. Han, M. Tong, L. Jin, W. Meng, A. Ren, Maternal age at pregnancy and risk for gestational diabetes mellitus among Chinese women with singleton pregnancies, Int. J. Diabetes Dev. Countries, 41 (2021), 114–120. https://doi.org/10.1007/s13410-020-00859-8 doi: 10.1007/s13410-020-00859-8
![]() |
[2] |
K. Zhang, H. Ma, Y. Zhao, H. Zan, L. Zhuang, The comparative experimental study of multilabel classification for diagnosis assistant based on Chinese obstetric EMRs, J. Healthcare Eng., (2018), 1–9. https://doi.org/10.1155/2018/7273451 doi: 10.1155/2018/7273451
![]() |
[3] |
C. Xu, P. Liu, Y. Sun, Research on disease prediction model for unbalanced medical datasets, Chin. J. Comput., 42 (2019), 596–609. https://doi.org/10.11897/SP.J.1016.2019.00596 doi: 10.11897/SP.J.1016.2019.00596
![]() |
[4] |
Y. Liu, H. Loh, A. Sun, Imbalanced text classification: A term weighting approach, Expert Syst. Appl., 36 (2009), 690–701. https://doi.org/10.1016/j.eswa.2007.10.042 doi: 10.1016/j.eswa.2007.10.042
![]() |
[5] | J. Stefanowski, Dealing with data difficulty factors while learning from imbalanced data, in Challenges in Computational Statistics and Data Mining, Springer, Cham, (2016), 333–363. https://doi.org/10.1007/978-3-319-18781-5_17 |
[6] | B. Zhou, Q. Cui, X. Wei, Z. Chen, BBN: Bilateral-branch network with cumulative learning for long-tailed visual recognition, in Proceedings of the IEEE/CVF Conference on Computer Vision and Pattern Recognition, (2020), 9719–9728. https://doi.org/10.1109/CVPR42600.2020.00974 |
[7] | B. Kang, S. Xie, M. Rohrbach, Z. Yan, A. Gordo, J. Feng, et al., Decoupling representation and classifier for long-tailed recognition, preprint, arXiv: 1910.09217. |
[8] |
Q. Yin, D. Shen, Y. Tang, Q. Ding, Intelligent monitoring of noxious stimulation during anaesthesia based on heart rate variability analysis, Comput. Biol. Med., 145 (2022), 105408. https://doi.org/10.1016/j.compbiomed.2022.105408 doi: 10.1016/j.compbiomed.2022.105408
![]() |
[9] |
T. Yan, P. Wong, C. Choi, C. Vong, H. Yu, Intelligent diagnosis of gastric intestinal metaplasia based on convolutional neural network and limited number of endoscopic images, Comput. Biol. Med., 126 (2020), 104026. https://doi.org/10.1016/j.compbiomed.2020.104026 doi: 10.1016/j.compbiomed.2020.104026
![]() |
[10] |
S. Wang, Y. Zhang, X. Cheng, X. Zhang, Y. Zhang, PSSPNN: PatchShuffle stochastic pooling neural network for an explainable diagnosis of COVID-19 with multiple-way data augmentation, Comput. Math. Methods Med., (2021), 1–18. https://doi.org/10.1155/2021/6633755 doi: 10.1155/2021/6633755
![]() |
[11] |
A. Rajkomar, E. Oren, K. Chen, A. Dai, N. Hajaj, M. Hardt, et al., Scalable and accurate deep learning with electronic health records, NPJ Digital Med., 1 (2018), 1–10. https://doi.org/10.1038/s41746-018-0029-1 doi: 10.1038/s41746-018-0029-1
![]() |
[12] |
A. Maxwell, R. Li, B. Yang, H. Weng, A. Ou, H. Hong, et al., Deep learning architectures for multi-label classification of intelligent health risk prediction, BMC. Bioinf., 18 (2017), 523. https://doi.org/10.1186/s12859-017-1898-z doi: 10.1186/s12859-017-1898-z
![]() |
[13] |
Z. Yang, Y. Huang, Y. Jiang, Y. Sun, Y. Zhang, P. Luo, et al., Clinical assistant diagnosis for electronic medical record based on convolutional neural network, Sci. Rep., 8 (2018), 6329. https://doi.org/10.1038/s41598-018-24389-w doi: 10.1038/s41598-018-24389-w
![]() |
[14] |
H. Liang, B. Tsui, H. Ni, C. Valentim, S. Baxter, G. Liu, et al., Evaluation and accurate diagnoses of pediatric diseases using artificial intelligence, Nat. Med., 25 (2019), 433. https://doi.org/10.1038/s41591-018-0335-9 doi: 10.1038/s41591-018-0335-9
![]() |
[15] |
N. Liu, E. Qi, M. Xu, B. Gao, G. Liu, A novel intelligent classification model for breast cancer diagnosis, Inf. Process. Manage., 56 (2019), 609–623. https://doi.org/10.1016/j.ipm.2018.10.014 doi: 10.1016/j.ipm.2018.10.014
![]() |
[16] |
C. Huang, X. Huang, Y. Fang, J. Xu, Y. Qu, P. Zhai, et al., Sample imbalance disease classification model based on association rule feature selection, Pattern Recognit. Lett., 133 (2020), 280–286. https://doi.org/10.1016/j.patrec.2020.03.016 doi: 10.1016/j.patrec.2020.03.016
![]() |
[17] |
B. Krawczyk, Learning from imbalanced data: Open challenges and future directions, Prog. Artif. Intell., 5 (2016), 221–232. https://doi.org/10.1007/s13748-016-0094-0 doi: 10.1007/s13748-016-0094-0
![]() |
[18] | X. Liu, X. Sun, Y. Meng, J. Liang, F. Wu, J. Li, Dice loss for data-imbalanced NLP tasks, in Proceedings of the 58th Annual Meeting of the Association for Computational Linguistics, (2020), 465–476. https://doi.org/10.48550/arXiv.1911.02855 |
[19] |
J. Yang, Z. Qu, Z. Liu, Improved feature-selection method considering the imbalance problem in text categorization, Sci. World J., (2014), 625342. https://doi.org/10.1155/2014/625342 doi: 10.1155/2014/625342
![]() |
[20] | F. Charte, A. Rivera, M. Jesus, F. Herrera, A first approach to deal with imbalance in multi-label datasets, in International Conference on Hybrid Artificial Intelligence Systems, (2013), 150–160. https://doi.org/10.1007/978-3-642-40846-5_16 |
[21] | F. Charte, A. Rivera, M. Jesus, F. Herrera, Concurrence among imbalanced labels and its influence on multilabel resampling algorithms, in International Conference on Hybrid Artificial Intelligence Systems, (2014), 110–121. https://doi.org/10.1007/978-3-319-07617-1_10 |
[22] | F. Charte, A. Rivera, M. Jesus, F. Herrera, Resampling multilabel datasets by decoupling highly imbalanced labels, in International Conference on Hybrid Artificial Intelligence Systems, (2015), 489–501. https://doi.org/10.1007/978-3-319-19644-2_41 |
[23] | X. Glorot, Y. Bengio, Understanding the difficulty of training deep feedforward neural networks, in Proceedings of the Thirteenth International Conference on Artificial Intelligence and Statistics, JMLR Workshop and Conference Proceedings, (2010), 249–256. http://proceedings.mlr.press/v9/glorot10a |
[24] |
F. Charte, A. Rivera, M. Jesus, F. Herrera, Addressing imbalance in multilabel classification: Measures and random resampling algorithms, Neurocomputing, 163 (2015), 3–16. https://doi.org/10.1016/j.neucom.2014.08.091 doi: 10.1016/j.neucom.2014.08.091
![]() |
[25] | D. Kingma, J. B. Adam, A method for stochastic optimization, preprint, arXiv: 1412.6980. |
[26] | P. Yang, X. Sun, W. Li, S. Ma, W. Wu, H. Wang, SGM: Sequence generation model for multi-label classification, in Proceedings of the 27th COLING, (2018), 3915–3926. https://doi.org/10.48550/arXiv.1806.04822 |
[27] | D. Lewis, Y. Yang, T. Rose, F. Li, Rcv1: A new benchmark collection for text categorization research, Mach. Learn. Res., (2004), 361–397. https://research.gold.ac.uk/id/eprint/29758 |
[28] |
M. Boutell, J. Luo, X. Shen, C. Brown, Learning multi-label scene classification, Pattern Recognit., 37 (2004), 1757–1771. https://doi.org/10.1016/j.patcog.2004.03.009 doi: 10.1016/j.patcog.2004.03.009
![]() |
[29] |
G. Tsoumakas, I. Katakis, Multi-label classification: An overview, Int. J. Data Warehous. Min., 3 (2007), 1–13. https://doi.org/10.4018/jdwm.2007070101 doi: 10.4018/jdwm.2007070101
![]() |
[30] | Y. Chen, Convolutional neural networks for sentence classification, in Proceedings of 2014 Conference on Empirical Methods in Natural Language Processing (EMNLP), (2014), 1–62. http://hdl.handle.net/10012/9592 |
1. | Cyrille Kenne, Pascal Zongo, René Dorville, A mathematical model for tilapia lake virus transmission with waning immunity, 2022, 16, 1751-3758, 98, 10.1080/17513758.2022.2033860 | |
2. | Cyrille Kenne, Gisèle Mophou, René Dorville, Pascal Zongo, A Model for Brucellosis Disease Incorporating Age of Infection and Waning Immunity, 2022, 10, 2227-7390, 670, 10.3390/math10040670 | |
3. | Flavius Guiaş, Equilibrium Solutions of a Modified SIR Model with Vaccination and Several Levels of Immunity, 2023, 18, 2224-2856, 550, 10.37394/23203.2023.18.57 | |
4. | Nursanti Anggriani, Lazarus Kalvein Beay, Meksianis Z. Ndii, Fatuh Inayaturohmat, Sanubari Tansah Tresna, A mathematical model for a disease outbreak considering waning-immunity class with nonlinear incidence and recovery rates, 2024, 6, 25889338, 170, 10.1016/j.jobb.2024.05.005 | |
5. | Liming Cai, Mingrui Lei, Jin Wang, Backward Bifurcation of an Age-Structured Epidemic Model with Partial Immunity, 2025, 35, 0218-1274, 10.1142/S0218127425500956 |