• Nem Talált Eredményt

Global stability of a multistrain SIS model with superinfection and patch structure

N/A
N/A
Protected

Academic year: 2022

Ossza meg "Global stability of a multistrain SIS model with superinfection and patch structure"

Copied!
10
0
0

Teljes szövegt

(1)

DOI: 10.1002/mma.6636

R E S E A R C H A R T I C L E

Global stability of a multistrain SIS model with superinfection and patch structure

Attila Dénes

1

Yoshiaki Muroya

2

Gergely Röst

1,3

1Bolyai Institute, University of Szeged, Szeged, Hungary

2Department of Mathematics, Waseda University, Tokyo, Japan

3Mathematical Institute, University of Oxford, Oxford, United Kingdom Correspondence

Attila Dénes, Bolyai Institute, University of Szeged, H-6720 Szeged, Hungary.

Email: denesa@math.u-szeged.hu Communicated by: R. Bravo de la Parra Funding information

EU-funded Hungarian, Grant/Award Number: EFOP-3.6.1-16-2016-00008;

János Bolyai Research Scholarship of the Hungarian Academy of Sciences; Ministry for Innovation and Technology,

Grant/Award Number:

TUDFO/47138-1/2019-ITM; National Research, Development and Innovation Fund of Hungary, Grant/Award Number:

FK 124016 and PD 128363

We study the global stability of a multistrain SIS model with superinfection and patch structure. We establish an iterative procedure to obtain a sequence of threshold parameters. By a repeated application of a result by Takeuchi et al. [Nonlinear Anal Real World Appl. 2006;7:235–247], we show that these parameters completely determine the global dynamics of the system: for any number of patches and strains with different infectivities, any subset of the strains can stably coexist depending on the particular choice of the parameters.

K E Y WO R D S

global asymptotic stability, multigroup epidemic model, multistrain model, patch model M S C C L A S S I F I C AT I O N

92D30; 34D23

1 I N T RO D U CT I O N

Several viruses have different genetic variants (subtypes) called strains which may differ in their infectivity and virulence.

Stronger strains might superinfect an individual already infected by another strain, and there can be a coexistence of different virus strains with different virulence. Nowak1 considered a model to provide an analytical understanding of the complexities introduced by superinfection. In our earlier work,2we considered a multistrain SIS model with super infection withninfectious strains and showed that it is possible to obtain a stable coexistence of any subgroup of then strains. We established an iterative method for calculating a sequence of reproduction numbers, which determine the strains being present in the globally asymptotically stable coexistence equilibrium.

Recently, there has been an increasing interest in the modelling of the spatial spread of infectious diseases (see, e.g., Arino and Portet,3Knipl,4Knipl and Röst,5Muroya, Kuniya and Enatsu,6Nakata and Röst7). There are several ways to model spatial spread: one might use partial differential equations (see, e.g., Peng and Zhao,8Allen et al.,9Ge et al.10) or one may apply ordinary or functional differential equations where individuals can travel between different patches (countries, regions, cities etc.).

Marvá et al.11considered a spatially distributed periodic multistrain SIS epidemic model with patches of periodic migra- tion rates without superinfection. Considering global reproduction numbers in the nonspatialized aggregated system that

This is an open access article under the terms of the Creative Commons Attribution License, which permits use, distribution and reproduction in any medium, provided the original work is properly cited.

© The Authors. Mathematical Methods in the Applied Sciences published by John Wiley & Sons Ltd

Math Meth Appl Sci. 2020;1–10. wileyonlinelibrary.com/journal/mma 1

(2)

serve to decide the eradication or endemicity of the epidemic in the initial spatially distributed nonautonomous model and comparing these global reproductive numbers with those corresponding to isolated patches, they showed that adequate periodic fast migrations can in many cases reverse local endemicity and get global eradication of the epidemic.

Motivated by our earlier work on multistrain models and by the recent results on spatial spread of diseases, we extend our previous model2to the general case ofppatches. In Section 2, we establish a multistrain SIS model with superinfection withninfectious strains and patch structure. In Section 3, we establish an iterative procedure to determine the globally asymptotically stable equilibrium of the multipatch model introduced in Section 2.

2 T H E M O D E L

We consider a heterogeneous virus population withnvirus strains having different infectivities and virulences. We will assume that superinfection is possible, and more virulent strains outcompete the less virulent ones in an infected indi- vidual taking over the host completely, that is, we assume that an infected individual is always infected by only one virus strain. Letndenote the number of strains with different virulences, whereaspstands for the number of patches. On each patch, the population is divided inton+1 compartments depending on the presence of any of the virus strains: the sus- ceptible class of patch𝓁is denoted byS𝓁(t)and on each patch𝓁, there areninfected compartmentsT1𝓁,,Tn𝓁where a larger index corresponds to a compartment of individuals infected by a strain with larger virulence, so fori<j,Tjindi- viduals superinfectTiindividuals. LetB𝓁denote the birth rate andb𝓁the death rate on the𝓁th patch. We denote by𝛽k𝑗𝓁 the transmission rate on patch𝓁by which thekth strain infects those who are infected by thejth strain. The transmission rates from susceptibles to strainkon patch𝓁will be denoted by𝛽kk𝓁. Recovery rate on patch𝓁among those infected by thekth strain will be denoted by𝜃𝓁k. Bym𝓁i, we denote the travel rate from patchito𝓁, which, on a given patch is equal for all compartments on that patch. This assumption—which is natural in the case of mild diseases—will be important in the transformation of variables described in the next section. The parametersB𝓁,b𝓁,m𝓁i,𝓁,i= 1,…,pare assumed to be nonnegative.

Using these notations, we consider the following multistrain SIS model with superinfection and patch structure:

dS𝓁(t)

dt =B𝓁b𝓁S𝓁(t) −S𝓁(t)

n

k=1

𝛽kk𝓁Tk𝓁(t) +

n

k=1

𝜃𝓁kT𝓁k(t) +

p

i=1

(1−𝛿𝓁i){

m𝓁iSi(t) −mi𝓁S𝓁(t)} , dT𝓁

k(t)

dt =S𝓁(t)𝛽kk𝓁Tk𝓁(t) +Tk𝓁(t)

n 𝑗=1

(1−𝛿k𝑗)𝛽k𝑗𝓁T𝓁𝑗(t) −(

b𝓁+𝜃k𝓁) T𝓁k(t) +

p

i=1

(1−𝛿𝓁i){

m𝓁iTki(t) −mi𝓁Tk𝓁(t)} , k=1,2,…,n, 𝓁=1,2,…,p,

(1)

with initial conditions

S𝓁(0) =𝜑𝓁0,Tk𝓁(0) =𝜑𝓁k,k=1,2,,n,𝓁=1,2,,p, (𝜑10, 𝜑11, 𝜑12,, 𝜑1n, 𝜑20, 𝜑21, 𝜑22,, 𝜑2n,, 𝜑p0, 𝜑p1, 𝜑p2,, 𝜑pn

)∈R(n+1)p+ =∶ Γ, (2)

where𝛿kjdenotes the Kronecker delta such that𝛿kj=1 ifk=jand𝛿kj=0 otherwise, and where 𝛽k𝓁𝑗=𝛽kk𝓁, 1≤𝑗k, and

𝛽k𝓁𝑗= −𝛽𝑗𝑗𝓁, k+1≤𝑗n, k=1,2,,n, 𝓁=1,2,,p. (3)

Note that forn=2 andp=1, (1) corresponds to the model by A. Dénes and G. Röst describing the spread of ectopar- asites and ectoparasite-borne diseases,12,13 whereas forp= 1, it corresponds to the multistrain SIS model by A. Dénes, Y. Muroya and G. Röst.2

3 M A I N R E S U LT

Let us introduce the notation

Nn𝓁(t) =S𝓁(t) +

n 𝑗=1

T𝑗𝓁(t),𝓁=1,2,…,p. (4)

(3)

Then, by (3), we have𝛽k𝓁𝑗= −𝛽𝑗𝓁kforkjand hence,

n

k=1

Tk𝓁(t)

n 𝑗=1

(1−𝛿k𝑗)𝛽k𝓁𝑗T𝑗𝓁(t) =0,𝓁=1,2,,p.

Thus, (1) is equivalent to dT𝓁k(t)

dt = (

Nn𝓁(t) −

n 𝑗=1

T𝑗𝓁(t) )

𝛽kk𝓁Tk𝓁(t) +Tk𝓁(t)

n 𝑗=1

(1−𝛿k𝑗)𝛽k𝓁𝑗T𝓁𝑗(t) −(

b𝓁+𝜃k𝓁) Tk𝓁(t)

+

p

i=1

(1−𝛿𝓁i){

m𝓁iTki(t) −mi𝓁Tk𝓁(t)}

, k=1,2,…,n−1,

(5a)

dTn𝓁(t) dt =

( Nn𝓁(t) −

n 𝑗=1

T𝓁𝑗(t) )

𝛽nn𝓁 Tn𝓁(t) +Tn𝓁(t)

n 𝑗=1

(1−𝛿n𝑗)𝛽n𝑗𝓁T𝓁𝑗(t)

−( b𝓁+𝜃𝓁n

)Tn𝓁(t) +

p

i=1

(1−𝛿𝓁i){

m𝓁iTin(t) −mi𝓁T𝓁n(t)}

=T𝓁n(t) (

𝛽nn𝓁 Nn𝓁(t) −

n 𝑗=1

{𝛽nn𝓁 − (1−𝛿n𝑗)𝛽n𝓁𝑗}

T𝑗𝓁(t) −( b𝓁+𝜃n𝓁

)) +

p

i=1

(1−𝛿𝓁i){

m𝓁iTni(t) −mi𝓁Tn𝓁(t)} ,

=T𝓁n(t)(

𝛽nn𝓁 Nn𝓁(t) −𝛽nn𝓁T𝓁n(t) −( b𝓁+𝜃n𝓁

))+

p

i=1

(1−𝛿𝓁i){

m𝓁iTin(t) −mi𝓁T𝓁n(t)} ,

(5b)

dNn𝓁(t)

dt =B𝓁b𝓁Nn𝓁(t) +

p

i=1

(1−𝛿𝓁i){

m𝓁iNni(t) −mi𝓁Nn𝓁(t)} , 𝓁=1,2,,p.

(5c)

Equations (5b) and (5c) are clearly independent from the rest of the equations. In particular, Equation (5c) are also independent from Equation (5b). As the coefficient matrixAof the linear system of equations

(B1

Bp

)

=

⎛⎜

⎜⎜

b1+∑p

i=1(1−𝛿1i)mi1 −m12 … −m1p

−m21 b2+∑p

i=1(1−𝛿2i)mi2 … −m2p

⋮ ⋮ ⋱ ⋮

−mp1 −mp2bp+∑p

i=1(1−𝛿pi)mip

⎞⎟

⎟⎟

⎠ (Nn1

Nnp

)

is a strictly diagonally dominantZ-matrix, it is nonsingular, and its inverse is nonnegative (because of the nonnegativity of the parameters), hence, this algebraic system has a unique, positive solution

(Nn1∗

Nnp∗

)

=A−1 (B1

Bp

) .

Let us defineP𝓁(t) ∶=Nn𝓁(t) −Nn𝓁∗, 𝓁=1,…,p, then forP𝓁(t), we have the equation d

dt (P1(t)

Pp(t)

)

= −A (P1(t)

Pp(t)

)

. (6)

From the properties of the matrix−A, applying the Gershgorin circle theorem, we obtain thatP𝓁(t)→0 exponentially ast→∞,𝓁=1,…,p. Hence, for Equation (5c), there exist positive constantsNn𝓁∗, 𝓁=1,2,…,psuch that

t→+∞lim Nn𝓁(t) =Nn𝓁∗, 𝓁=1,2,…,p, (7)

(4)

exponentially and (5b) has the following limit system:

dT𝓁n(t)

dt =Tn𝓁(t)(

𝛽nn𝓁 Nn𝓁∗−( b𝓁+𝜃𝓁n

)−𝛽nn𝓁 T𝓁n(t)) +

p

i=1

(1−𝛿𝓁i){

m𝓁iTni(t) −mi𝓁Tn𝓁(t)}

, 𝓁=1,2,,p, (8)

which is ap-dimensional Lotka–Volterra system with patch structure, in the form as Equation (2.1) in Takeuchi et al.14 We introduce the notation

̃ mii=

p 𝓁=1

(1−𝛿i𝓁)mi𝓁, i=1,2,,p,

and define the connectivity matrix

M=

⎡⎢

⎢⎢

11 m12m1p m2122m2p

⋮ ⋮ ⋱ ⋮

mp1 mp2 … −pp

⎤⎥

⎥⎥

.

Now, we define

c𝓁n=𝛽nn𝓁 Nn𝓁∗− (b𝓁+𝜃n𝓁), 𝓁=1,2,,p,

and

Mn=

⎡⎢

⎢⎢

c1n11 m12m1p

m21 c2n22m2p

⋮ ⋮ ⋱ ⋮

mp1 mp2cpnpp

⎤⎥

⎥⎥

.

Let us denote bys(L)the stability modulus of ap×pmatrixL, defined bys(L) ∶=max{Re𝜆∶𝜆is an eigenvalue ofL}. If Lhas nonnegative off-diagonal elements and is irreducible, thens(L)is a simple eigenvalue ofLwith a (componentwise) positive eigenvector (see, e.g., Theorem A.5 in Smith15).

Proposition 1 (see Theorem 2.1 in Takeuchi et al.14).Suppose that Mnis irreducible. Then, Equation (8) has a positive equilibrium which is globally asymptotically stable if s(Mn)>0. If s(Mn) ≤ 0, then 0 is a globally asymptotically stable equilibrium, and the populations go extinct in every patch.

Note that we may take that the populations go extinct in every patch not only ifs(Mn) < 0 but also ifs(Mn) = 0 (see Theorem 2.2 of Faria16).

Let En = (Tn1∗,T2∗n ,,Tnp∗) be the unique equilibrium of (8) which is globally asymptotically stable. Then,En = (0,0,,0)ifs(Mn) ≤ 0, andEn = (Tn1∗,T2∗n ,,Tnp∗)satisfiesTn𝓁∗ >0, 𝓁 =1,2,,p, ifs(Mn) >0. Therefore, in the first case, the unique equilibrium of (8), is globally asymptotically stable on{(Tn1,Tn2,,Tnp) ∈Rp+}, whereas in the sec- ond case, the unique positive equilibriumEn= (Tn1∗,Tn2∗,,Tp∗n )withTn𝓁∗>0, 𝓁=1,2,,pis globally asymptotically stable with respect to{(Tn1,Tn2,,Tnp) ∈Rp+}⧵{(0,0,…,0)}. Let us introduce the notations

Nn−1𝓁 (t) =S𝓁(t) +

n−1

𝑗=1

T𝑗𝓁(t), 𝓁=1,2,,p,

and

b𝓁(1) =b𝓁𝛽kn𝓁Tn𝓁∗ =b𝓁+𝛽nn𝓁 Tn𝓁∗, k=1,2,,n−1, 𝓁=1,2,,p,

(5)

and

B𝓁(1) =B𝓁+𝜃n𝓁Tn𝓁∗, 𝓁=1,2,,p,

where(Tn1∗,,Tn1∗)is either equal to(0,,0)(ifs(Mn) ≤ 0) or it is equal to the unique positive equilibrium of (8) (if s(Mn)>0). This way, substitutingTni∗, 1 =1,,pinto the place ofTin(t)in (4) and (5), we may consider the following reduced system of (5) for the global stability of (1):

dT𝓁

k(t) dt =

(

Nn−1𝓁 (t) −

n−1 𝑗=1

T𝑗𝓁(t) )

𝛽kk𝓁Tk𝓁(t) +Tk𝓁(t)

n−1

𝑗=1

(1−𝛿k𝑗)𝛽k𝑗𝓁T𝑗𝓁(t) − (

b𝓁(1)+𝜃k𝓁) Tk𝓁(t)

+

p

i=1

(1−𝛿𝓁i){

m𝓁iTki(t) −mi𝓁Tk𝓁(t)}

,k=1,2,,n−2,

(9a)

dTn−1𝓁 (t)

dt =

(

Nn−1𝓁 (t) −

n−1 𝑗=1

T𝓁𝑗(t) )

𝛽n−1,n−1𝓁 T𝓁n−1(t) +Tn−1𝓁 (t)

n−1 𝑗=1

(1−𝛿n−1,𝑗)𝛽n−1,𝑗𝓁 T𝑗𝓁(t) − (

b𝓁(1)+𝜃n−1𝓁 ) Tn−1𝓁 (t)

+

p

i=1

(1−𝛿𝓁i){

m𝓁iTin−1(t) −mi𝓁Tn−1𝓁 (t)}

=Tn−1𝓁 (t)

(𝛽n−1,n−1𝓁 Nn−1𝓁 (t) −𝛽n−1,n−1𝓁 Tn−1𝓁 (t) − (

b𝓁(1)+𝜃n−1𝓁 )) +

p

i=1

(1−𝛿𝓁i){

m𝓁iTin−1(t) −mi𝓁Tn−1𝓁 (t)} ,

(9b) dNn−1𝓁 (t)

dt =B𝓁(1)b𝓁(1)Nn−1𝓁 (t) +

p

i=1

(1−𝛿𝓁i){

m𝓁iNn−1i (t) −mi𝓁Nn−1𝓁 (t)} , 𝓁=1,2,,p.

(9c)

It is easy to see that (9) is of similar structure as (5), but with dimensionpn. The positivity of the new parameters fol- lows from the conditions (3). This means that by repeating the above steps, namely, substituting the limit of the total populations in the patches and then substituting the limit of the Lotka–Volterra system for the strongest strain, we can further reduce the dimension by substituting the values of the equilibrium which is globally asymptotically stable, of the decoupledpdimensional Lotka–Volterra system into the remaining equations.

We proceed repeating the same steps for the newly arising reduced system, decreasing the dimension of our system in each round of the procedure byp. In each round, forqdecreasing fromn−1 to 1, we introduce the respective limits Nq𝓁∗andTq𝓁∗, as well as the matricesMqcorresponding to the reduced system in an analogous way as it was presented in the case of the original system. In the end, we arrive at apdimensional Lotka–Volterra system, the dynamics of which can be determined in a similar way as in the above case. This final system will give us an equilibrium value forS1(t)and (T11(t),T12(t),…,Tp1(t)). Thus, by the above discussion, we can reach a conclusion by induction to the global dynamics of the model (1) and we formulate the following theorem.

Theorem 1. Assume that the connectivity matrix M is irreducible. Then the global dynamics of the multistrain, mul- tipatch SIS model (1) is completely determined by the threshold parameters(s(M1),s(M2),…,s(Mn))which can be obtained iteratively. There exists an equilibrium inΓwhich is globally asymptotically stable with respect to the regionΓ0, whereΓ0is the interior ofΓ.

Proof of Theorem 1. The main part of the proof consists of the above description of the steps of the procedure. There is one point left to be shown: we have to prove that in each step, when we substitute the limitsN𝜅𝓁∗andT𝜅𝓁∗, respectively, into the remaining equations, the dynamics of the resulting system is indeed equivalent to that of the preceding one.

We summarize the steps of the procedure in the following.

1. We obtainNn𝓁∗ (𝓁=1,,p)from the linear system (6).

2. We substitute the limitsNn𝓁 (𝓁=1,…,p)into Equation (5b) to obtain Equation (8).

(6)

3. We obtain the limitsTn𝓁∗(𝓁=1,…,p)of the Lotka–Volterra system (8).

4. We create the new variablesNn−1𝓁 (t), 𝓁=1,,pand parametersb𝓁(1),B𝓁(1),𝓁=1,,p.

5. We substitute the limitsTn𝓁 (𝓁 = 1,,p)into Equation (5a) to obtain the reduced system (9) which has the same structure as the original one (5).

6. We repeat this cyclen−1 times, with the indices decreased by 1 every time.

For the validity of Step 3 in the qth cycle, we need to verify that Mn−q is irreducible. Because Mn−q = M + diag[c1n−q,,cpn−q]and we assumed thatMis irreducible,Mn−qis also irreducible.

To obtain that in each case, the limit of the solutions of the resulting system after the substitution will be the same equilibrium as the limit of the solutions of the original system, we will apply Theorem 4.1 of Hirsch and Smith.17To apply this theorem, we recall the quasimonotone condition17for a differential equationx(t) =f(t,x(t)): we say that the time-dependent vector field𝑓J×D→Rn(whereJ RandD Rn) satisfies the quasimonotone condition inDif for all(t,y),(t,z) ∈J×D, we have

𝑦zand𝑦i=ziimplies𝑓i(t, 𝑦)≤𝑓i(t,z).

According to Theorem 4.1 of Hirsch and Smith,17 if𝑓,gJ×D→ Rnare continuous, Lipschitz on each compact subset ofD, at least one of them satisfies the quasimonotone condition, andf(t,y)g(t,y)for all(t,y) ∈J×D, then

𝑦,z∈Rn, 𝑦zimpliesx(t;t0, 𝑦)x(t;t0,z)for allt>t0, wherex(t;t0,y)denotes the solution ofx(t) =f(t,x(t))started fromyatt=t0.

To show that the limitsT𝜅𝓁∗ obtained during the procedure by substituting the limits of (8) into (5a) are the same as the limit of the variablesT𝜅𝓁,𝜅 =1,…,n,𝓁=1,…,pin the original system, we will use an induction argument.

It is clear from the above that the claim is true for𝜅 =n. Let now 1rn−1 and let us suppose that the claim holds for allT𝜅𝓁(t)forr< 𝜅n. The limitsTr𝓁 are obtained by first substituting the limitsTr+1𝓁 into the equations forT𝑗𝓁(t), 1 ≤ jrand then substituting the limitsNr𝓁∗into the equations forTr𝓁(t), hence, we have to compare the limits of the two systems

dTr𝓁(t) dt =(

Nr+1𝓁 (t) −2Tr+1𝓁 (t) −Tr𝓁(t))

𝛽rr𝓁Tr𝓁(t) − (

b𝓁(n−r+1)+𝜃r𝓁

) T𝓁r(t) +

p

i=1

(1−𝛿𝓁i){

m𝓁iTri(t) −mi𝓁Tr𝓁(t)}

=(

Nr𝓁(t) −Tr+1𝓁 (t) −Tr𝓁(t))

𝛽rr𝓁T𝓁r(t) − (

b𝓁(n−r+1)+𝜃r𝓁

) Tr𝓁(t) +

p

i=1

(1−𝛿𝓁i){

m𝓁iTri(t) −mi𝓁T𝓁r(t)} (10)

and

dTr𝓁(t) dt =(

Nr𝓁∗Tr𝓁(t))

𝛽rr𝓁Tr𝓁(t) − (

b𝓁(n−r)+𝜃r𝓁

) T𝓁r(t) +

p 𝓁=1

(1−𝛿𝓁i){

m𝓁iTri(t) −mi𝓁Tr𝓁(t)} , 𝓁=1,2,,p.

(11)

We know thatNr𝓁(t)(𝓁=1,,p) converge toNr𝓁(𝓁=1,,p), whereas from the definition ofr, we have that Tr+1𝓁 (t)(𝓁=1,…,p) converge toT𝓁∗r+1(𝓁=1,…,p). Then, for any𝜀 >0, there exists āt>0 such that|Nr𝓁(t)−Nr𝓁∗|<

𝜀and|Tr+1𝓁 (t) −Tr+1𝓁 |< 𝜀for allt > ̄t,𝓁 = 1,,p. If we substituteTr+11∗ +𝜀,,Tr+1p∗ +𝜀,Nr1∗𝜀,,Nn−qp∗𝜀, resp.Tr+11∗𝜀,,Tr+1p∗𝜀,Nr1∗+𝜀,,Nn−qp∗ +𝜀into (10), we obtain two systems of the same structure as (11), and one of them is a lower, the other is an upper estimate of (10), and each has a globally asymptotically stable equilibrium (T1r(𝜀),…,Tpr(𝜀))

and(T1r(𝜀),…,Tpr(𝜀)), respctively, because of Proposition 1. It is easy to see that the original system (10), considered as a nonautonomous system with time-dependent coefficientsTr+11 (t),,Tr+1p (t),Nr1(t),,Nrp(t), satisfies the quasimonotone condition, as well as the systems obtained after the substitution. Hence, we can apply Theorem 4.1 of Hirsch and Smith17to obtain that for any solution(Tr1(t),…,Trp(t))of (10),

T𝓁r(𝜀)≤lim inf

t→∞ Tr𝓁(t)≤lim sup

t

T𝓁r(t)≤T𝓁r(𝜀), 𝓁=1,…,p. (12)

(7)

Solutions of limit Eq. (11) converge to a globally asymptotically stable equilibrium by Proposition 1, and by letting 𝜀→0, we find that this limit is the same as that of (10).

As we have assumed that for all larger indices, the limits of the compartments of the original system (5) are equal to the limits obtained during the procedure, using the equations forT1r(t),,Trp(t)afternr+1 cycles of the procedure satisfy the quasimonotone condition and the comparison (12), the limits obtained for these have to coincide with those of the original system (forr=n, the statement follows directly).

To prove that not only attractivity but also global asymptotic stability holds, we will again use induction. LetE= (1,T11,,T1n,, ̄Sp,Tp1,,Tpn)denote the equilibrium obtained at the end of the procedure, whereT𝑗i = 0 or T𝑗i > 0 depending on the stability moduli(s(M1),s(M2),…,s(Mn))and letE𝜅 = (1,T11,,T1𝜅,, ̄Sp,Tp1,,Tp𝜅) be the equilibrium of thep(𝜅+1)-dimensional system

dTk𝓁(t) dt =

( N𝜅𝓁(t) −

𝜅 𝑗=1

T𝓁𝑗(t) )

𝛽kk𝓁T𝓁k(t) +T𝓁k(t)

𝜅 𝑗=1

(1−𝛿k𝑗)𝛽k𝓁𝑗T𝑗𝓁(t) − (

b𝓁(n−𝜅)+𝜃k𝓁) Tk𝓁(t)

+

p

i=1

(1−𝛿𝓁i){

m𝓁iTki(t) −mi𝓁T𝓁k(t)} , k=1,2,, 𝜅−1, 𝓁=1,2,,p,

(13a) dT𝜅𝓁(t)

dt =T𝜅𝓁(t)

(𝛽𝜅,𝜅𝓁 N𝜅𝓁(t) − (

b𝓁(n−𝜅)+𝜃𝓁𝜅)

𝛽𝜅,𝜅𝓁 T𝜅𝓁(t) )

+

p

i=1

(1−𝛿𝓁i){

m𝓁iT𝜅i(t) −mi𝓁T𝓁𝜅(t)} , 𝓁=1,2,,p,

(13b)

dN𝜅𝓁(t)

dt =B𝓁(n−𝜅)b𝓁(n−𝜅)N𝜅𝓁(t) +

p

i=1

(1−𝛿𝓁i){

m𝓁iN𝜅i(t) −mi𝓁N𝜅𝓁(t)} , 𝓁=1,2,,p,

(13c)

obtained during the procedure withE𝜅consisting of the firstp(𝜅+1)coordinates ofE. Let us suppose thatE𝜅is a stable equilibrium of thep(𝜅+1)-dimensional reduced system for some𝜅n. We will show that in each step,E𝜅+1

is a stable equilibrium of thep(𝜅+2)-dimensional reduced system dT𝓁

k(t) dt =

(

N𝜅+1𝓁 (t) −

𝜅+1

𝑗=1

T𝑗𝓁(t) )

𝛽kk𝓁Tk𝓁(t) +Tk𝓁(t)

𝜅+1

𝑗=1

(1−𝛿k𝑗)𝛽k𝑗𝓁T𝑗𝓁(t)

− (

b𝓁(n−𝜅−1)+𝜃k𝓁) Tk𝓁(t) +

p

i=1

(1−𝛿𝓁i){

m𝓁iTki(t) −mi𝓁T𝓁k(t)} , k=1,2,…, 𝜅,

(14a)

dT𝜅+1𝓁 (t)

dt =

(

N𝜅𝓁+1(t) −

𝜅+1

𝑗=1

T𝑗𝓁(t) )

𝛽𝜅𝓁+1,𝜅+1T𝓁𝜅+1(t) +T𝓁𝜅+1(t)

𝜅+1

𝑗=1

(1−𝛿𝜅+1,𝑗)𝛽𝜅𝓁+1,𝑗T𝑗𝓁(t)

− (

b𝓁(n−𝜅−1)+𝜃𝜅+1𝓁 )

T𝜅+1𝓁 (t) +

p

i=1

(1−𝛿𝓁i){

m𝓁iT𝜅+1i (t) −mi𝓁T𝓁𝜅+1(t)} ,

=T𝜅𝓁+1(t)

(𝛽𝜅𝓁+1,𝜅+1N𝜅𝓁+1(t) −𝛽𝜅𝓁+1,𝜅+1T𝜅𝓁+1(t) − (

b𝓁(n−𝜅−1)+𝜃𝜅𝓁+1)) +

p

i=1

(1−𝛿𝓁i){

m𝓁iTi𝜅+1(t) −mi𝓁T𝜅𝓁+1(t)} ,

(14b)

dN𝜅+1𝓁 (t)

dt =B𝓁(n−𝜅−1)b𝓁(n−𝜅−1)N𝜅+1𝓁 (t) +

p

i=1

(1−𝛿𝓁i){

m𝓁iN𝜅+1i (t) −mi𝓁N𝜅+1𝓁 (t)} , 𝓁=1,2,…,p.

(14c)

(8)

Suppose this does not hold, that is,E𝜅+1 is unstable. In this case, there exist an 𝜀 > 0 and a sequence {xm} → E𝜅+1, |xmE𝜅+1| < 1∕m such that the orbits started from the points of the sequence leave B(E𝜅+1, 𝜀) ∶=

{

x∈R(𝜅+2)p+ |xE𝜅+1|≤𝜀}

. By an exit point fromB(E𝜅+1, 𝜀), we mean a pointxsuch that|E𝜅+1x| = 𝜀and for the trajectory throughx, there is an open intervalJ ∋ 0 such that for alltJ,xtB(E𝜅+1, 𝜀)ift ≤ 0 and xtB(E𝜅+1, 𝜀)ift>0. Let us denote byx𝜀mthe first exit point fromB(E𝜅+1, 𝜀)of the solution started fromxm, reached at time𝜏m. There is a convergent subsequence of the sequencex𝜀m(still denoted byxm𝜀) which tends to a point denoted byx𝜀S(E𝜅+1, 𝜀) ∶={

x∈R(𝜅+2)p+ ∶|xE𝜅+1|=𝜀}

. We will show thatE𝜅+1𝛼(x𝜀). For this end, let us consider the setS

(

E𝜅+1,𝜀2)

. Clearly, all solutions started from the pointsxm(we drop the first elements of the sequence, if necessary) will leave the setB

(

E𝜅+1,𝜀2)

. We denote the last exit point of each trajectory from this set before time𝜏m, respectively, byxm𝜀∕2. Also this sequence has a convergent subsequence (still denoted the same way), let us denote its limit byx𝜀∕2 . We will show that the trajectory started fromx𝜀∕2 goes throughx𝜀. AsE𝜅+1is globally attractive, this tra- jectory will eventually enterS

(

E𝜅+1,4𝜀)

at some timeT>0. Let us suppose that the trajectory started fromx𝜀∕2does not go throughx𝜀and let us denote byd>0 the distance of this trajectory fromx𝜀. For continuity reasons, there is an N∈Nso that for anym>N,|x𝜀∕2tx𝜀∕2m t|<max

{d 2,𝜀8}

for 0<t<T. This means that formlarge enough, the tra- jectory started fromx𝜀∕2m will enter againS

(

E𝜅+1,𝜀2)

without getting close tox𝜀which contradicts eitherx𝜀mbeing the first exit point fromB(E𝜅+1, 𝜀)orx𝜀m∕2being the last exit point before𝜏mfromB

(

E𝜅+1,𝜀2)

. Hence, we have shown that the trajectory started fromx𝜀∕2goes throughx𝜀. Proceeding like this (taking neighbourhoods of radius𝜀∕4, 𝜀∕8 etc.), we obtain that the backward trajectory ofx𝜀enters any small neighbourhood ofE𝜅+1. That is, there exists a decreas- ing sequencetn < 0 such that|xtnE𝜅+1| < 2n𝜀. We have eithertnt* for somet* < 0 ortn → −∞. In the first case,xtnxt* = E𝜅+1which contradicts the fact thatE𝜅+1is an equilibrium. Hence,tn → ∞, and we obtain that E𝜅+1𝛼(x𝜀), while it follows from the global attractivity ofE𝜅+1that the𝜔-limit set of the trajectory is{E𝜅+1}. Let us denote this trajectory by𝛾(x𝜀)

We know that Equations (14b) for d

dtT𝜅1+1(t),…,dtdT𝜅p+1(t)and (14c) for d

dtN𝜅1+1(t),…,dtdN𝜅p+1(t)can be decoupled from the rest of the equations and using the exponential stability of the limits

t→+∞lim N𝜅𝓁+1(t) =N𝜅𝓁∗+1,𝓁=1,2,,p,

and Proposition 1 we obtain thatT1𝜅+1,,Tp𝜅+1is a stable equilibrium of the system consisting of the system dT𝜅+1𝓁 (t)

dt =T𝜅𝓁+1(t)(

𝛽𝜅𝓁+1,𝜅+1N𝜅𝓁∗+1−(

b𝓁+𝜃𝜅𝓁+1)

𝛽𝜅𝓁+1,𝜅+1T𝜅𝓁+1(t)) +

p

i=1

(1−𝛿𝓁i){

m𝓁iT𝜅i+1(t) −mi𝓁T𝓁𝜅+1(t)}

,𝓁=1,2,,p.

Therefore, the equilibriumE𝜅+1is stable in the coordinatesT𝜅+11 ,,T𝜅+1p in the sense that for anỹ𝜀 >0 there exists a ̃𝛿(̃𝜀) >0 such that for any initial valuexwith|x−E𝜅+1| < ̃𝛿,|T𝜅+1𝓁 (t) −T𝓁𝜅+1| < ̃𝜀for allt > 0 and𝓁 = 1,…,p.

Thus, the trajectory𝛾(x𝜀)obtained above lies entirely in the subspace {

T𝜅+11 =T1𝜅+1,,Tp𝜅+1=Tp𝜅+1 }

. On the other hand, the currentp(𝜅+2)-dimensional system coincides with thep(𝜅+1)-dimensional system on this subspace.

For the latter system, stability of the equilibriumE𝜅follows from the induction assumption. However, the existence of an orbit̃𝛾 (different from the equilibriumE𝜅+1) whose𝜔-limit set is{E𝜅+1}and whose𝛼-limit set containsE𝜅+1 contradicts the stability of the equilibriumE𝜅. Indeed, let us suppose thatE𝜅is stable and there exists such an orbit

̃𝛾. The stability ofE𝜅would implies that for any ̂𝜀 >0, there exists a ̂𝛿(̂𝜀)such that for any solution started from an initial valuePwith|P−E𝜅|< ̂𝛿, we have|Pt−E𝜅|< ̂𝜀for allt>0. Hence, this is also true for̂𝜀=|E𝜅P|∕2 for anỹ ̃𝛾, which is a contradiction, as a solution started from a point of̃𝛾clearly leavesS(E𝜅, ̂𝜀). Hence, no such orbit ̂𝛾 can exist. This implies the global asymptotic stability of the equilibrium of thep(𝜅+2)-dimensional system.

Hivatkozások

KAPCSOLÓDÓ DOKUMENTUMOK

We consider the case when the two regions are symmetric in the population sizes and values of every model parameter, but have different local reproduction numbers, and we assume that

We present distributed algorithms for effectively calculating basic statistics of data using the recently introduced newscast model of computation and we demonstrate how to

We present a novel approach to study the local and global stability of fam- ilies of one-dimensional discrete dynamical systems, which is especially suitable for difference

In addition, the latest literature [8, 21,36] and [6,10–12, 14,35] have not touched the global attractivity of the positive equilibrium point for the patch structure neoclassical

In the present study we measure the fluorescence kinetics of NADH in an aqueous solution with high precision and apply a custom machine learning based analysis method to

- the method of the investigation of the self-excitation and that of the stability of the machine having an elliptic field is in its method iden- tical with the analysis with

We establish an iterative method for calculating a sequence of reproduction numbers which determine what strains are present in the globally asymptotically stable

We prove that in the case of bidirectional transportation on two different patches, the system has threshold dynamics: the disease free equilib- rium is globally asymptotically