Abstract
To discuss the impact factors on the spread of infectious diseases, we study a free boundary problem describing a SIS (susceptible-infected-susceptible) model in a heterogeneous environment. Firstly, the existence and uniqueness of the global solution are given. Then the basic reproduction number related to time is defined, and a spreading-vanishing dichotomy of infectious diseases is obtained. The impacts of the diffusion rate of infected individuals, expanding capability, and the scope and scale of initial infection on the spreading and vanishing of infectious disease are analyzed. Numerical simulations are given to show that the large expanding capability is unfavorable to the prevention and control of the disease.
Keywords: SIS model, Free boundary, Risk index, Spreading and vanishing
MSC: primary, 35K57, 92D30, secondary, 35K55
1. Introduction
Infectious diseases are still the major causes of suffering and death in developing countries and developed countries (Hethcote, 2000), and usually have a significant impact on population size and historical events (McNeill, 1976). For example, the West Nile virus caused an encephalitis outbreak in New York City in 1999 (Center for Disease Control and Prevention (CDC), 1999). According to the latest real-time statistics from the World Health Organization (WHO), COVID-19, as of May 2022, has infected more than 516 million individuals, and at least 6.2 million people have died worldwide. However, it is impracticable to study the spreading of infectious diseases by experiments. To understand the spread and control of infectious diseases, mathematical models as an important tool have been attracting much attention (Abdelrazec, Belair, Shan, & Zhu, 2016; Hethcote, 2000; Keeling & Rohani, 2008; Martcheva, 2015; Thomas & Urena, 2001).
To discuss the impact of spatial heterogeneity of environment and movement of individuals on infectious diseases, an SIS epidemic reaction-diffusion system was proposed by Allen et al. in (Allen, Bolker, Lou, & Nevai, 2008), and the giving system is
(1.1) |
where S(x, t) and I(x, t) denote the density of susceptible and infected individuals at location x and time t, respectively. The positive constants dS and dI account for the diffusion rates of susceptible and infected individuals, the positive bounded Hlder continuous functions β(x) and γ(x) can be interpreted as rates of disease transmission and recovery for x ∈ Ω, respectively.
The authors in (Allen et al., 2008) characterized the risk of the region by the basic reproduction number . The DFE (disease-free equilibrium) is always unstable and there exists a unique EE (endemic equilibrium) in the high-risk domain , and the DFE is stable if and only if infected individuals have mobility above a threshold value for the low-risk domain . In addition, letting N = S + I, adding two equations in (1.1) and then integrating over Ω conclude that for t > 0, , which implies that the total population size remains unchanged. More work on system (1.1) with constant total population size can be found in the literature (Cui & Lou, 2016; Peng, 2009; Peng & Liu, 2009; Peng & Zhao, 2012).
Considering changeable total population size, Li, Peng and Wang (Li, Peng, & Wang, 2017) studied the following SIS epidemic with a linear external source
(1.2) |
where dS, dI, β, γ, S and I have the same epidemiological interpretations as in (1.1). The linear function Λ − S represents the external source (or supply) for the susceptible population, for more explanation, please refer to literature (Gao & Ruan, 2011; Hethcote, 2000). The authors in (Li et al., 2017) mainly discussed the global stability of the unique endemic equilibrium when spatial environment is homogeneous, and the asymptotic profile of endemic equilibria if the diffusion rate of the susceptible or infected population is small or large.
Previous studies have focused on SIS epidemic models in a fixed domain. In the real world, the change of biological habitat is caused by the movement of species, which can be described mathematically by free boundary. In fact, the free boundary is widely used in many fields; for example, a single population model with a free boundary for invasive species was proposed by Du and Lin in (Du & Lin, 2010), which was extended to the competition model in (Cao, Li, Wang, & zhao, 2021; Du, Wang, & Zhou, 2017; Du & Lin, 2014; Wang & Zhao, 2014), the predator-prey model in (Lin, 2007; Takhirov & Norov, 2019; Wang & Zhao, 2017; Yousefnezhad, Mohammadi, & Bozorgnia, 2018), and a mutualistic model with advection in (Li & Lin, 2015). In recent years, a free boundary has also been introduced into epidemic models. For example, an SIR epidemic model with free boundary was discussed in (Huang & Wang, 2015; Kim, Lin, & Zhang, 2013; Zhu, Guo, & Lin, 2017), and a SIRS model with free boundary was considered in (Cao et al., 2017a). Ge et al. in (Ge, Kim, Lin, & Zhu, 2015) studied a simplified SIS model with advection and free boundary and gave the spreading speeds when spreading happens; a diffusion-advection simplified SIS epidemic model in a heterogeneous time-periodic environment was studied in (Ge, Lei, & Lin, 2017); a SIS reaction-diffusion model with free boundary and mass action mechanism was discussed in (Wang & Guo, 2019); recently an SIS epidemic reaction-diffusion model with mass-action incidence incorporating spontaneous infection in a spatially heterogeneous environment was examined (Tong, Ahn, & Lin, 2021); a SIS model risk-induced dispersal of infected individuals has been studied (Choi, Lin, & Ahn, 2022); the spreading or vanishing of an SIS epidemic model with free boundary and nonlocal incidence rate was analyzed in (Cao et al., 2017b; Huang & Wang, 2019).
Based on above discussion, in this paper, we will consider the following SIS reaction-diffusion problem
(1.3) |
where the interval (g(t), h(t)) is unknown and will be given with (S, I). g′(t) = −μIx(g(t), t) (or h′(t) = −μIx(h(t), t)) is a special case of the well-known Stefan condition, which has been established in (Lin, 2007). Biologically, this system implies that outside of the free boundaries, there is no infectious, only susceptible individuals. The interval (g(t), h(t)), which depends on time t, is the habitat for the infectious individuals. Moreover, we suppose that initial functions S0 and I0 satisfy
(1.4) |
Further, we assume that
which implies that the contact rate (β) and the recovery rate (γ) are similar at the faraway places, and the faraway is high-risk.
It is worth mentioning that in the fixed domain, the authors in (Li et al., 2017) primarily discussed the asymptotic profile of endemic equilibria when the diffusion rate is small or large. However, we discuss how the diffusion coefficient affects the spreading and vanishing of disease described mathematically by the free boundary problem in this paper.
The organization of this article is as below. The existence and uniqueness of the global solution are obtained in Section 2. Section 3 is devoted to the definition and properties of the basic reproduction number. Section 4 establishes a spreading-vanishing dichotomy for problem (1.3). Some sufficient conditions for the disease to spread or vanish are given in Section 5. Finally, we explain our theoretical results by numerical simulations.
2. Global existence and uniqueness
At first, we discuss the existence and uniqueness of the local solution of problem (1.3), and then we show that the solution is global by suitable estimates.
Theorem 2.1
For any given initial value (S0, I0) satisfying (1.4), and any ν ∈ (0, 1), there exists a T > 0 such that problem (1.3) admits a unique solution (S, I; g, h) satisfying
(2.1) |
and
(2.2) |
Furthermore,
(2.3) |
where C and T only depend on , and .
Proof: Applying similar methods as in (Cao et al., 2017b; Chen & Friedman, 2000; Du & Lin, 2010), for system (1.3), we first straighten the free boundaries. Let ξ be a function in C3([0, ∞)) satisfying
Consider the transformation (y, t) → (x, t) such that
It is obvious to know that the above transformation x → y is a diffeomorphism from (−∞, + ∞) onto (−∞, + ∞) while and . Hence, it translates the left free boundary x = g(t) to the fixed line y = −h0 and the right free boundary x = h(t) to the fixed line y = h0. Directing calculations give
Let
then the free boundary problem (1.3) can be rewritten as
(2.4) |
where A = A(g(t), h(t), y), B = B(g(t), h(t), y) and C = C(g(t), h(t), y). The following proof first finds a suitable complete metric space, and then uses the contraction mapping theorem together with standard Lp theory and the Sobolev imbedding theorem (Ladyzenskaja, Solonnikov, & Ural’ceva, 1968) to obtain the existence and uniqueness of the solution for system (2.4). We are going to omit it here, see Theorem 2.1 in (Du & Lin, 2014) or (Li & Lin, 2015) for more details. □
Theorem 2.2
Let (S, I; g, h) be a bounded solution of problem (1.3) for t ∈ [0, T] with some T ∈ (0, + ∞), then there exist positive constants C1 and C2 independent of T such that
Proof: For any bounded solution (S, I; g, h) of problem (1.3) for t ∈ [0, T], using the initial value condition together with the strong maximum principle yields
and
Applying the Hopf's boundary lemma to the equation of I gives that
Therefore g′(t) < 0 and h′(t) > 0 for t ∈ (0, T] by the free boundary condition.
Let N = S + (1 + ϵ0)I with , direct computation yields
where .
Recalling that Ix(h(t)−, t) < 0 = Ix(h(t)+, t) and Ix(g(t)−, t) = 0 < Ix(g(t)+, t) for 0 < t ≤ T, we have
(2.5) |
Consider the following auxiliary problem
(2.6) |
where . For l > max{h(T), − g(T)}, taking for x ∈ [ − l, l] and t ∈ [0, T], , which satisfies
We claim that
(2.7) |
Otherwise, we suppose that , there exists a point (x0, t0) ∈ [ − l, l] × [0, T] so that min Wl = Wl(x0, t0) < 0. Due to , we know that t0 ≠ 0. Moreover, Wl(±l, t) > 0 for t ∈ [0, T], so x0 ≠ ± l. Therefore we have (x0, t0) ∈ (−l, l) × (0, T]. If (x0, t0) ∈ (−l, l) × (0, T] with x0 ≠ g(t0) and x0 ≠ h(t0), we have
which is a contradiction with (Λ∗ − Λ) − m(N∗ − N) > 0.
Assume that x0 = g(t0) or h(t0), without loss of generality, we suppose that x0 = g(t0). Due to I > 0 in (g(t0), h(t0)), and I = 0 in , we derive
that is
which immediately gives
it is a contradiction by Nx(g(t)+, t) > Nx(g(t)−, t). Hence, (2.7) holds. Let l → +∞, we have
Therefore
it follows from I(x, t) ≤ N(x, t) that I(x, t) ≤ C1 for x ∈ (g(t), h(t)) and t ∈ [0, T].
Finally, we show that − g′(t) ≤ C2 and h′(t) ≤ C2 for t ∈ (0, T]. The proof is analogous to that of Lemma 2.2 in (Du & Lin, 2010) with C2 = 2MμC1 and
So we omit it here. □
Theorem 2.3
System (1.3) admits a unique solution (S, I; g, h) for all t ∈ (0, + ∞).
Proof: Combining all estimates in Theorem 2.2 with the standard continuous extension method, we can prove that the unique solution (S, I; g, h) of system (1.3) exists for all t ∈ (0, + ∞). The proof is similar to Theorem 2.4 in (Cao et al., 2017b) or Theorem 2.3 in (Du & Lin, 2010), we omit it here. □
3. The basic reproduction number
In general, the basic reproduction number can be used to describe the dynamics of the temporal and spatial spread of the disease. We firstly discuss the basic reproduction number and its properties for the following problem
(3.1) |
which is given by linearizing the corresponding system of (1.3) in a bounded domain Ω at the disease-free equilibrium.
It is well-known (Allen et al., 2008; Ge et al., 2015) that the basic reproduction number for (3.1) is given by the variational formula
Due to above definition of the basic reproduction number, the following result was given in (Allen et al., 2008; Ge et al., 2015) and (Huang, Han, & Liu, 2010).
Proposition 3.1
((Ge et al., 2015)) Letλ0is the principal eigenvalue of the principle problem
then has the same sign as λ0.
Proposition 3.2
((Ge et al., 2015)) The following assertions hold:
is a positive and monotone decreasing function ofd;
asd → 0;
asd → ∞;
(iv) There exists a threshold value d∗ ∈ [0, + ∞) such that for d < d∗ and for d > d∗. If all sites in the domain are lower-risk (β(x) < γ(x) for x ∈ Ω), then for all d > 0;
(v) Let Bh be a ball in with the radius h. The is strictly monotone increasing function of h, that is if h1 < h2, then . Moreover, provided that (H) holds;
(vi) Assume that β∗ and γ∗ are positive constants. If Ω = (−h0, h0), β(x) ≡ β∗ and γ(x) ≡ γ∗, then
Pay attention to the interval (g(t), h(t)) of the free boundary problem (1.3) is changing with t, so the basic reproduction number is not a constant, and it varies with time t. Now we define the basic reproduction number for the free boundary problem (1.3) by
(3.2) |
which is called the spatial-temporal risk index in (Lin & Zhu, 2017) or the spatial-temporal basic reproduction number in (Zhu, Ren, & Zhu, 2018). We can define by the monotonicity in Theorem 3.3.
It follows from that Propositions 3.1 and 3.2, we have
Theorem 3.3
is strictly monotone increasing function oft, in other words,ift1 < t2;
(b) If g(t) → g∞ > −∞ and h(t) → h∞ < ∞ as t → +∞, we have that ;
(c) If h(t) − g(t) → ∞ as t → +∞, then provided (H) holds.
Proof: (a) − g′(t) > 0 and h′(t) > 0 for t ∈ (0, + ∞) in Theorem 2.2 imply that − g(t) and h(t) are strictly increasing functions. It is easy to see that is strictly monotone increasing function of t by (v) of Proposition 3.2.
(b) According to g∞ > −∞ and h∞ < ∞, we consider the problem
So the basic reproduction number for above problem is given by the variational formula
From (v) of Proposition 3.2 and monotonicity of g(t) and h(t), then holds for any t > 0, i.e.,
It follows that
(3.3) |
On the other hand, duo to the definition of , for any given ε > 0, there exists a function such that and
By definition of , for given ε > 0, there exists a function such that .
Note that
Let hε = dist(suppψ, ∂(g∞, h∞)), which means that ψ(x) = 0 in (g∞, g∞ + hε) and (h∞ − hε, h∞). In view of g(t) → g∞ and h(t) → h∞ as t → +∞, there exists a large T > 0 such that for t > T,
and
where positive constant M only depends on and . According to the monotonicity of g(t) and h(t) yields
Hence,
In light of the arbitrariness of ε, we can obtain
(3.4) |
which together with (3.3) gives that (b) holds.
(c) Suppose that h(t) → +∞ as t → +∞. For any given ε > 0, there exists a large and positive constant l such that
for x > l provided that (H) holds. Moreover, there exists T0 > 0 such that h(t) > l for all t ≥ T0. For each fixed t ∈ [T0, ∞), let Φ(x) ∈ C2(g(t), h(t)) satisfy
Due to the definition of , for t ∈ [T0, + ∞), we have
Hence
In view of the arbitrariness of ε, we can obtain
If g(t) → −∞ as t → +∞, we can draw a conclusion that from the above discussion.
Hence, holds provided that (H) holds and h(t) − g(t) → +∞ as t → +∞.
□
Remark 3.1
If g(t) → −∞ or h(t) → +∞ as t → +∞, then provided that (H) holds.
4. Spreading-vanishing dichotomy
In this section, we are going to establish a spreading-vanishing dichotomy for the free boundary problem (1.3). We will prove that if the interval (g(t), h(t)) is finite in the end, then I → 0 uniformly as t → +∞. Rather, for each solution (S, I) of problem (1.3), if the infected individuals spread to the whole area, that is, (g(t), h(t)) → (−∞, + ∞) as t → +∞.
From Theorem 2.2, we obtain that g′(t) < 0 and h′(t) > 0 for t ∈ (0, + ∞). Hence g(t) is monotonically decreasing and h(t) is monotonically increasing with t. So their limits exist, i.e.,
where g∞ ∈ [ − ∞, − h0) and h∞ ∈ (h0, + ∞].
Theorem 4.1
If − ∞ < g∞ < h∞ < + ∞, then , and locally uniformly for , where satisfies
(4.1) |
Proof: At first, we show that . We argue indirectly, and assume that
Then there exits a sequence {(xk, tk)} with tk ∈ (0, ∞) and xk ∈ [g(tk), h(tk)] such that
for all , and tk → +∞ as k → +∞.
Due to − ∞ < g∞ < h∞ < + ∞, then xk ∈ [g(t), h(t)] ⊆ [g∞, h∞], there exists a subsequence of {(xk, tk)}, denoted by {(xk, tk)}, such that xk → x0 as k → +∞, where x0 ∈ [g∞, h∞].
Define
for x ∈ [g(t + tk), h(t + tk)] and t ∈ [ − tk, + ∞). Using Theorem 2.1 and the standard parabolic regularity, there exists a subsequence of {(Sk, Ik)} such that
where satisfies
In view of , then in (g∞, h∞) × (0, + ∞) by the maximum principle.
Note that , it follows from the Hopf's boundary lemma that
Hence for all large enough ki, we have
and
On the other hand, in view of − ∞ < g∞ < h∞ < + ∞, we have obtain
which is a contradiction. Therefore .
Since , for any ε > 0, there exists a sufficiently large T > 0 such that
(4.2) |
Now, for any given 0 < L < + ∞, in view of (4.2), we have for (x, t) ∈ [ − L, L] × [T, + ∞), where satisfies
(4.3) |
and satisfies
(4.4) |
It is well-known that
where and are unique positive steady state of problems (4.4) and (4.3), respectively. It is easily seen that
(4.5) |
by the standard parabolic regularity, where S∗(x) is the unique solution to problem (4.1).
Therefore, we obtain that
uniformly for x ∈ [ − L, L]. According to (4.5), we have
The arbitrariness of L implies that
Theorem 4.2
Ifh∞ − g∞ = +∞, then .
Proof: Assume that by contradiction. It follows from the proof of Theorem 4.1 that
(4.6) |
Since h∞ − g∞ = +∞, from the Remark 3.1, there exists a large enough T∗ > 0 such that
In view of (4.6), for any , there exists a T∗∗ > T∗ such that
Following, we consider the eigenvalue problem
it has the principal eigenvalue λ0 < 0 by Proposition 3.1 and the corresponding eigenvalue function ψ(x) with .
Obviously, I(x, t) satisfies
We consider following auxiliary problem
(4.7) |
and construct a suitable lower solution for problem (4.7).
Define
where δ > 0 is small enough such that δψ(x) ≤ I(x, T∗∗) for x ∈ [g(T∗), h(T∗)].
For t > T∗∗ and g(T∗) < x < h(T∗), direct computation yields
provided that .
Due to the upper and lower solutions method and comparison principle, we obtain
Hence we have that , which is a contradiction.
Next, we show that the left and right of free boundary are finite or infinite simultaneously in the case where β and γ are positive constants.
Theorem 4.3
Assume thatβandγare positive constants. Ifg∞ > −∞ or h∞ < + ∞, then − ∞ < g∞ < h∞ < + ∞.
Proof: Without loss of generality, we suppose that h∞ < + ∞. Assume that g∞ = −∞ by a contradiction, then it follows from Theorem 3.1 and (H) that and there exists a large T1 > 0 such that . So for t > T1. In view of Proposition 3.1, we have λ1 < 0, where λ1 is a principal eigenvalue of the following problem
and φ1(x) is corresponding eigenfunction. Hence, there exists a ε > 0 such that
has principal eigenvalue and eigenfunction ψ1(x) > 0 with in (g(T1), h(T1)). In addition,
Obviously, using parabolic comparison principle we have S(x, t) ≥ Sm for , where .
From above result, it is easy to see that
And the following problem
(4.8) |
admits a unique positive solution satisfying in (g(T1), h(T1)). Using the Hopf's boundary lemma yields ω′(g(T1)) > 0 > ω′(h(T1)). So ω′(x) has at least one zero point in (g(T1), h(T1)). Let the largest zero of ω′(x) in (g(T1), h(T1)) be x0. Then
Next, we consider the following auxiliary problem
(4.9) |
In order to obtain for t ≥ T1, we will show that ε1φ1(x) is a lower solution of problem (4.9) in (g(T1), h(T1)) × (T1, + ∞). Direct calculation gives
provided that . Meanwhile, , . If ε1 is sufficiently small, then
Hence, we can apply the comparison principle to conclude that
if ε1 is sufficiently small with . Therefore,
which implies that has a positive lower bound for t ≥ T1.
Finally, based on the above results, we construct a suitable lower solution of I(x, t) in (x0, h(t)) × (T1, + ∞).
Set
and calculate
Using the fact that h′(t) → 0 and h(t) → h∞ < + ∞ as t → +∞, we can find T2 > T1 such that and then
since that ω′ ≤ 0 and for (x, t) ∈ [x0, h(t)] × [T2, + ∞]. Since and , we have
for (x, t) ∈ [x0, h(t)] × [T2, + ∞].
In addition, we choose ε2 small such that ε2ω(x0) ≤ ε1φ1(x0) and
Applying the comparison principle yields
It follows that
as t → +∞. By the free boundary condition, we have
as t → +∞, which contradicts with the assumption h∞ < + ∞. Hence, − ∞ < g∞ < h∞ < + ∞ holds.
According to Theorems 4.1 and 4.2, we can obtain the following spreading-vanishing dichotomy:
Theorem 4.4
Let (S, I; g, h) be the solution to problem (1.3). Then, the following alternatives hold:
(i) Spreading: h∞ − g∞ = +∞ and then ; or
(ii) Vanishing: h∞ − g∞ < + ∞ and then , and locally uniformly for , where S∗(x) satisfies problem (4.1).
Remark 4.1
Theorem 4.3 implies that h∞ = −g∞ = ∞ when spreading happens if β and γ are positive constants.
5. Sufficient condition for vanishing or spreading
First, recalling the definition of , we have the following result from the proof of Theorem 4.2.
Theorem 5.1
Ifh∞ − g∞ < + ∞, then . Moreover, if for any t0 ≥ 0, then − g∞ = h∞ = +∞.
It is well-known in (Allen et al., 2008; Li et al., 2017; Peng, 2009; Peng & Liu, 2009; Peng & Zhao, 2012) that can be used as a threshold to describe the spreading or vanishing of infectious diseases on fixed boundary, that is, if , then infectious diseases is spreading; rather, the infectious diseases is vanishing. For the free boundary problem (1.3), we will show that spreading happens when ; the condition can not ensure for any t > 0. We will show that the spreading and vanishing of infectious disease are possible when , which mainly depends on the variables μ and I0.
Theorem 5.2
Suppose, then
if is small enough.
Proof: We prove it by constructing a suitable upper solution for I(x, t), which is similar to that of Lemma 5.3 in (Du & Lin, 2010) or (Ge et al., 2015). Due to , we consider the following eigenvalue problem
which has a positive solution ψ(x) with , and principal eigenvalue λ0 > 0. In view of Theorem 2.2, we know that S(x, t) ≤ C1 for . Hence,
Set
and
Clearly W(−η(t), t) = 0 and W(η(t), t) = 0. Direct computations yield that
and
Hence − η′(t) ≤ −μWx(−η(t), t) and η′(t) ≥−μWx(η(t), t) hold if ε is sufficiently small. And then, from the definition of η(t) we know that − η(0) < g(0) = −h0 < h(0) = h0 < η(0). On the other hand, by direct calculations, for t > 0 and x ∈ (−η(t), η(t)), we obtain
provided that δ is sufficiently small and satisfies .
Finally, we choose small enough such that . Applying the comparison principle ((Du & Lin, 2010)) for free boundary problem yields
That is
So
if is sufficiently small.
From the above proof, taking and , we can prove similarly that ω(x, t) is still the upper solution of I(x, t) for (x, t) ∈ (g(t), h(t)) × [0, + ∞) if positive constant M is sufficiently large and μ is small enough. Therefore, the following result holds.
Theorem 5.3
Suppose, then
if μ is sufficiently small.
When , Theorem 5.3 indicates that vanishing happens for small expanding capability μ, the following result shows that spreading occurs for large expanding capability μ.
Theorem 5.4
Assume, thenh∞ − g∞ = +∞ if μ is sufficiently large and (H) holds.
Proof: According to (v) of Proposition 3.2, we have
if (H) holds, and then there exists a sufficiently large positive constant M > 0 so that . Set
Using the Lemma 2.5 in (Zhu & Lin, 2018) we can conclude that there exists a μM > 0 such that
for μ > μM.
On the other hand, due to the monotonicity of gμ(t) and hμ(t), there exists a T0 > 0 such that
Moreover,
Hence, if μ > μM, then h∞ − g∞ = +∞ holds by Theorem 5.1. □
Due to Theorems 5.3 and 5.4, the following result holds, see details in Theorem 5.5 in (Ge et al., 2015).
Theorem 5.5
Fixh0andI0, there existsμ∗ ∈ [0, + ∞) such that if μ > μ∗, then spreading happens; and if 0 < μ ≤ μ∗, then vanishing happens.
Finally, we will discuss how the diffusion coefficient d affects the vanishing and spreading of the disease.
Theorem 5.6
If the setG = {x ∈ [ − h0, h0], β(x) > γ(x)} ≠ ∅, then there exists d∗ > 0 such that − g∞ = h∞ = +∞ for d ≤ d∗ or d > d∗ and μ is large enough.
Proof: Due to Proposition 3.2 (ii) and G ≠ ∅ we conclude that
There is a positive constant d∗ such that , and for d ≤ d∗ by Proposition 3.2 (i) − (iii). Theorem 5.1 implies that − g∞ = h∞ = +∞ for d ≤ d∗, while if d > d∗, then Theorem 5.4 gives that − g∞ = h∞ = +∞ provided μ is large enough. □
Similarly, based on Proposition 3.2 (ii) and (iii), and Theorems 5.2 and 5.3, the following result for vanishing holds.
Theorem 5.7
Fixh0 > 0, β(x) and γ(x). There exists a nonnegative constant such that − ∞ < g∞ < h∞ < + ∞ for provided that μ or is small enough. If β(x) < γ(x) in [ − h0, h0], then , that is, vanishing happens if μ or is small enough.
Proof: It is easily seen that there is a constant so that for by Proposition 3.2 (i) and (iii). If β(x) < γ(x) in [ − h0, h0], then by Proposition 3.2 (i) and (ii).
Since , then vanishing happens if μ or is small enough by Theorems 5.2 and 5.3. □
6. Simulation and discussion
At first, we use numerical simulations to explain impact expanding of the capability μ on spreading of infectious diseases. Taking some functions as follows:
we can know that β±∞ = 3.6, γ±∞ = 3.55 and (H) holds. Moreover, according to (3.2), we have
Next, we discuss the asymptotic behaviors of the solution to problem (1.3) and the changing of free boundaries by choosing different expanding capabilities.
Example 6.1
Taking μ = 50. Theorems 5.3 and 5.5 show that the solution I decays to zero if μ is small. It is easy Fig. 1 that the disease tends to extinction quickly, and the free boundaries expand slowly.
Example 6.2
Letting μ = 60. It follows from Theorems 5.4 and 5.5 that the spreading of the disease happens for large μ. It is easily seen that a spatially inhomogeneous stationary endemic state appears for large μ. The free boundaries expand quickly.
Fig. 1.
μ = 50. The left graph shows that the solution I decays to zero quickly. The right graph is the corresponding planar graph, which shows the free boundaries expand slowly and will be limited in a long run.
In this paper, we mainly discuss a SIS model with a linear source and variable total population, and the free boundaries are used to describe the spreading left and right fronts. Firstly, we present the local existence and uniqueness of solution for problem (1.3) by the contraction mapping theorem together with standard Lp theory and the Sobolev embedding theorem (see Theorem 2.1), and then we prove that the solution is global solution by some estimates (see Theorems 2.2 and 2.3). Next, the basic reproduction number is defined for the SIS diffusion-reaction system with a free boundary. And then, a spreading-vanishing dichotomy for problem (1.3) is given (see Theorem 4.4). We then prove that is used as the threshold to describe the spreading or vanishing of the disease. If for any t0 ≥ 0, then the disease is spreading (see Theorems 4.2 and 5.1). When , whether the disease is spreading or vanishing depends on the initial value of the infected individuals I0 and the expanding capability μ. That is to say, if , the disease is vanishing when initial value of the infected individuals I0 is small enough (Theorem 5.2) or the expanding capability μ is small enough (Theorem 5.3); the disease is spreading when the expanding capability μ is large enough and (H) holds (Theorem 5.4). Finally, we deal with the impact of the coefficient d for the disease. There exists a d∗ > 0, for d ≤ d∗ or d > d∗ and μ is sufficiently large, then the disease spreads (Theorem 5.6). On the other hand, there exists a , the disease vanishing provided if μ or I0 are sufficiently small (Theorem 5.7). If the diffusion rates dS and dI are different, we believe that complicated dynamical behaviors would appear (Cui & Lou, 2016; Peng, 2009; Peng & Zhao, 2012).
In addition, our simulations indicate that expanding capability μ plays an important role in the spreading and vanishing of disease. If , the disease will vanish when the expanding capability μ is comparatively small (Fig. 1), and the disease will spread when the expanding capability μ is comparatively large (Fig. 2).
Fig. 2.
μ = 60. The solution I stabilizes to an equilibrium in the left graph. The right planar graph shows that the free boundaries expand fast.
When spreading occurs, the spreading speed for the simplified SIS model was given in (Ge et al., 2015). Owing to the coupled system for our reaction-diffusion model, some technical and mathematical difficulties for the spreading speed deserve further study.
Declaration of competing interest
We declare that we have no financial and personal relationships with other people or organizations that can inappropriately influence our work, there is no professional or other personal interest of any nature or kind in any product, service and/or company that could be construed as influencing the position presented in, or the review of, the manuscript entitled, “The impact factors of the risk index and diffusive dynamics of a SIS free boundary model”.
Handling Editor: Dr HE DAIHAI HE
Footnotes
The first author is supported by the Postgraduate Research & Practice Innovation Program of Jiangsu Province (KYCX21-3188), the second author is supported under the framework of international cooperation program managed by the National Research Foundation of Korea (NRF-2019K2A9A2A06025237) and the third author is supported by the National Natural Science Foundation of China (Grant No. 12271470).
Peer review under responsibility of KeAi Communications Co., Ltd.
References
- Abdelrazec A., Belair J., Shan C.H., Zhu H.P. Modeling the spread and control of dengue with limited public health resources. Mathematical Bioscience. 2016;271:136–145. doi: 10.1016/j.mbs.2015.11.004. [DOI] [PubMed] [Google Scholar]
- Allen L.J.S., Bolker B.M., Lou Y., Nevai A.L. Asymptotic profiles of the steady states for an SIS epidemic reaction-diffusion model. Discrete Contin. Dyn. Syst. 2008;21:1–20. [Google Scholar]
- Cao J.F., Li W.T., Wang J., Yang F.Y. A free boundary problem of a diffusive SIRS model with nonlinear incidence. Zeitschrift für Angewandte Mathematik und Physik. 2017;68:1–16. [Google Scholar]
- J. F. Cao, W. T. Li, J. Wang, M. zhao, The dynamics of a Lotka-Volterra competition model with nonlocal diffusion and free boundaries, Advances in Differential Equations, 26 (2021), 163-200.
- Cao J.F., Li W.T., Yang F.Y. Dynamics of a nonlocal SIS epidemic model with free boundary. Discrete and Continuous Dynamical Systems - Series B. 2017;22:247–266. [Google Scholar]
- Center for Disease Control and Prevention (CDC) Update: West Nile-like viral encephalitis-New York, 1999. Morbidity & Mortality Weekly Report. 1999;48:890–892. [PubMed] [Google Scholar]
- Chen X.F., Friedman A. A free boundary problem arising in a model of wound healing. SIAM Journal on Mathematical Analysis. 2000;32:778–800. [Google Scholar]
- W. Choi, Z. G. Lin, I. Ahn, SIS reaction-diffusion model with risk-induced dispersal under free boundary, Nonlinear Analysis: Real World Applications 21 (2022), 103605.
- Cui R.H., Lou Y. A spatial SIS model in advective heterogeneous environments. Journal of Differential Equations. 2016;261:3305–3343. [Google Scholar]
- Du Y.H., Lin Z.G. Spreading-Vanishing dichotomy in the diffusive logistic model with a free boundary. SIAM Journal on Mathematical Analysis. 2010;42:377–405. [Google Scholar]
- Du Y.H., Lin Z.G. The diffusive competition model with a free boundary: Invasion of a superior or inferior competitor. Discrete and Continuous Dynamical Systems - Series B. 2014;19:3105–3132. [Google Scholar]
- Du Y.H., Wang M.X., Zhou M.L. Semi-wave and spreading speed for the diffusive competition model with a free boundary. Journal de Mathematiques Pures et Appliquees. 2017;9:253–287. [Google Scholar]
- Gao D.Z., Ruan S.G. An SIS patch model with variable transmission coefficients. Mathematical Biosciences. 2011;232:110–115. doi: 10.1016/j.mbs.2011.05.001. [DOI] [PubMed] [Google Scholar]
- Ge J., Kim K.I., Lin Z.G., Zhu H.P. A SIS reaction-diffusion-advection model in a low-risk and high-risk domain. Journal of Differential Equations. 2015;259:5486–5509. [Google Scholar]
- Ge J., Lei C.X., Lin Z.G. Reproduction numbers and the expanding fronts for a diffusion-advection SIS model in heterogeneous time-periodic environment. Nonlinear Analysis: Real World Applications. 2017;33:100–120. [Google Scholar]
- Hethcote H.W. The mathematics of infectious diseases. SIAM Review. 2000;42:599–653. [Google Scholar]
- Huang W.Z., Han M.A., Liu K.Y. Dynamics of an SIS reaction-diffusion epidemic model for disease transmission. Mathematical Biosciences and Engineering. 2010;7:51–66. doi: 10.3934/mbe.2010.7.51. [DOI] [PubMed] [Google Scholar]
- Huang H.M., Wang M.X. The reaction-diffusion system for an SIR epidemic model with a free boundary. Discrete and Continuous Dynamical Systems - Series B. 2015;20:2039–2050. [Google Scholar]
- Huang H.M., Wang M.X. A nonlocal SIS epidemic problems with double free boundaries. Zeitschrift für Angewandte Mathematik und Physik. 2019;70:1–19. [Google Scholar]
- Keeling M.J., Rohani P. Princeton University Press; New Jersey: 2008. Modeling infectious disease in humans and animals. [Google Scholar]
- Kim K.I., Lin Z.G., Zhang Q.Y. An SIR epidemic model with free boundary. Nonlinear Analysis: Real World Applications. 2013;14:1992–2001. [Google Scholar]
- Ladyzenskaja O.A., Solonnikov V.A., Ural’ceva N.N. Amer. Math. Soc; Providence, RI: 1968. Linear and quasilinear equations of parabolic type. [Google Scholar]
- Li M., Lin Z.G. The spreading fronts in a mutualistic model with advection. Discrete and Continuous Dynamical Systems - Series B. 2015;20:2089–2105. [Google Scholar]
- Lin Z.G. A free boundary problem for a predator-prey model. Nonlinearity. 2007;20:1883–1892. [Google Scholar]
- Lin Z.G., Zhu H.P. Spatial spreading model and dynamics of West Nile virus in birds and mosquitoes with free boundary. Journal of Mathematical Biology. 2017;75:1381–1409. doi: 10.1007/s00285-017-1124-7. [DOI] [PubMed] [Google Scholar]
- Li H.C., Peng R., Wang F.B. Varying total population enhances disease persistence: Qualitative analysis on a diffusive SIS epidemic model. Journal of Differential Equations. 2017;262:885–913. [Google Scholar]
- Martcheva M. Springer; New York: 2015. An introduction of mathematical epudemiology. [Google Scholar]
- McNeill W.H. Blackwell; Oxford: 1976. Plagues and people. [Google Scholar]
- Peng R. Asymptotic profiles of the positive steady state for an SIS epidemic reaction-diffusion model. I. Journal of Differential Equations. 2009;247:1096–1119. [Google Scholar]
- Peng R., Liu S.Q. Global stability of the steady states of an SIS epidemic reaction- diffusion model. Nonliner Analysis. 2009;71:239–247. [Google Scholar]
- Peng R., Zhao X.Q. A reaction-diffusion SIS epidemic model in a time-periodic environment. Nonlinerity. 2012;25:1451–1471. [Google Scholar]
- Takhirov J.O., Norov A. On a predator-prey model with free boundary. Uzbek Math. J. 2019;4:162–168. [Google Scholar]
- Thomas D., Urena B. A reaction-diffusion SIS epidemic model in a time-periodic environment. Mathematical and Computer Modelling. 2001;34:771–781. [Google Scholar]
- Y. Tong, I. Ahn, Z. G. Lin, Effect of diffusion in a spatial SIS epidemic model with spontaneous infection, Discrete and Continuous Dynamical Systems - Series B 34 (2021), 4045-4057.
- Wang Y.Z., Guo S.J. A SIS reaction-diffusion model with a free boundary condition and nonhomogeneous coefficients. Discrete and Continuous Dynamical Systems - Series B. 2019;24:1627–1652. [Google Scholar]
- Wang M.X., Zhao J.F. Free boundary problems for a Lotka-Volterra competition system. Journal of Dynamics and Differential Equations. 2014;26:655–672. [Google Scholar]
- Wang M.X., Zhao J.F. A free boundary problem for the predator-prey model with double free boundaries. Journal of Dynamics and Differential Equations. 2017;29:957–979. [Google Scholar]
- Yousefnezhad M., Mohammadi S.A., Bozorgnia F. A free boundary problem for a predator-prey model with nonlinear prey-taxis. Applied Mathematics. 2018;63:125–147. [Google Scholar]
- Zhu M., Guo X.F., Lin Z.G. The risk index for an SIR epidemic model and spatial spreading of the infectious disease. Mathematical Biosciences and Engineering. 2017;14:1565–1583. doi: 10.3934/mbe.2017081. [DOI] [PubMed] [Google Scholar]
- Zhu M., Lin Z.G. The impact of human activity on the risk index and spatial spreading of dengue fever. Nonlinear Analysis: Real World Applications. 2018;39:424–450. [Google Scholar]
- Zhu D.D., Ren J.L., Zhu H.P. Spatial-temporal basic reproduction number and dynamics for a dengue disease diffusion model. Mathematical Methods in the Applied Sciences. 2018;41:5388–5403. [Google Scholar]