<<

www.nature.com/scientificreports

OPEN Dynamics of the stochastic chemostat with Monod-Haldane response function Received: 14 June 2017 Liang Wang1,2, Daqing Jiang1,3,5, Gail S. K. Wolkowicz2 & Donal O’Regan4,5 Accepted: 20 September 2017 The stochastic chemostat model with Monod-Haldane response function is perturbed by environmental Published: xx xx xxxx white noise. This model has a global positive solution. We demonstrate that there is a stationary distribution of the stochastic model and the system is ergodic under appropriate conditions, on the basis of Khasminskii’s theory on ergodicity. Sufcient criteria for extinction of the microbial population in the stochastic system are established. These conditions depend strongly on the Brownian motion. We fnd that even small scale white noise can promote the survival of populations, while large scale noise can lead to extinction. Numerical simulations are carried out to illustrate our theoretical results.

Te chemostat, known as a continuous stir tank reactor (CSTR) in the engineering literature, is a basic piece of apparatus used for the continuous culture of . It occupies a central place in mathemat- ical and has played an important role in many felds. In ecology it is ofen viewed as a model of simple lake or an ocean system. It can also model the wastewater treatment process1 or biological waste decomposition2. In some processes, the chemostat plays a central role in the commercial production of genetically altered organisms. Te theoretical investigation of the dynamics of microbial interaction in the chemostat was initiated by Monod3 and Novick et al.4 in 1950. Tere is also an extensive literature on the chemostat (for exam- ple, see refs5–11 and the references therein for recent research) concerned with the dynamics of various types of chemostat models. Te classic chemostat model with single species and single limiting substrate takes the form

 dS()t 1  =−QS((0 St)) − ((St))xt(),  δ  dt  dx()t  =−μ((St))xt() Qx()t ,  dt (1.1) where S(t), x(t) stand for the concentrations of nutrient and microbial population at time t respectively; S0 denotes the imput concentration of nutrient and Q represents the volumetric fow rate of the mixture of nutrient and microorganism; the coefcient δ is the ratio of the biomass of the microbial population produced by the nutrient consumed. Te growth rate of the microbial population is represented by the function μ(S) (μ()Sm≤<∞), which is generally assumed to be non-negative. Tat is μ(0)0= , μ(S) > 0 for S > 0. Some experiments and observations indicate that not only insufcient nutrient but also excessive nutrient may inhibit the growth of a microbial population in the chemostat. To model such growth, Andrews12 suggested a non-monotonic response function: mS μ()S = , aS++KS2 (1.2)

1School of Mathematics and Statistics, Northeast Normal University, Changchun, 130024, P. R. China. 2Department of Mathematics and Statistics, McMaster University, Hamilton, ON L8S 4K1, Canada. 3College of Science, China University of Petroleum (East China), Qingdao, 266580, P. R. China. 4School of Mathematics, Statistics and Applied Mathematics, National University of Ireland, Galway, Ireland. 5Nonlinear Analysis and Applied Mathematics (NAAM)- Research Group, King Abdulaziz University, Jeddah, Saudi Arabia. Correspondence and requests for materials should be addressed to D.J. (email: [email protected])

SCientifiC RepOrTs | 7:13641 | DOI:10.1038/s41598-017-13294-3 1 www.nature.com/scientificreports/

which is called the Monod-Haldane growth rate (inhibition rate). Parameter m is the maximum growth rate of the microorganism, a is the Michaelis-Menten (or half-saturation) constant. Te term KS2 models the inhibitory efect of the substrate at high concentrations. Te chemostat model with Monod-Haldane response function has been studied by many researchers (see e.g. Wang et al.13, Pang et al.14, Baek et al.15 and the references there in). Taking the non-monotone functional response into account, the chemostat model (1.1) becomes:

dS 0 1 mS  =−QS()S − x,  dt δ aS++KS2  dx mS  = −.  2 xQx  dt aS++KS (1.3) System (1.3) has a trivial equilibrium point (S0, 0). It is an asymptotically stable steady node when mS0 mS0 ⁎⁎ Q > 2 and a saddle point when Q < 2 . In addition, there are two interior equilibria (,Sx11), aS++00KS aS++00KS ⁎⁎ ⁎ ⁎ (,Sx22) and S1 , S2 satisfy the equation QKSQ2 +−()mS+=Qa 0, Qm<. (1.4) It is not difcult to verify that one of the two interior equilibria is a stable nodal point and the other is a saddle point. In mathematical biology, when the orbits tend to the stable nodal point (S*, x*), the continuous culture of microorganism is considered to be successful. We refer the reader to Chen16 for more details. System (1.3) is deterministic with parameters assumed to be constant. Environmental fuctuations are ignored, which, from the biological point of view, is unrealistic. Ecosystem dynamics are inevitably afected by environ- mental noise and it is more realistic to include the efect of stochasticity rather than to study models that are entirely deterministic. With respect to chemostat models, even though the experimental results that are observed in well-controlled laboratory conditions have been shown to match closely with the prediction of deterministic models involving ordinary diferential equations, we cannot ignore the diference that may occur in operational conditions. Imhof et al.17 derived a stochastic chemostat model by considering a discrete-time Markov process with jumps corresponding to the addition of a centered Gaussian term to the deterministic model. With the time step converging to zero, they reported that the stochastic model may lead to extinction in some cases in which the deterministic model predicts persistence. In18, Campillo et al. considered a set of stochastic chemostat models that are valid on diferent scales. Xu and Yuan19 dealt with the stochastic chemostat in which the maximal growth ∼ rate m is perturbed by white noise and obtained a new break-even concentration λ which completely determines the persistence or extinction of the microbial population. Zhao and Yuan20 further computed the probability for extinction and persistence in mean of the microbial population in19 using stochastic calculus. Wang et al.21 inves- tigated the periodic solutions for the stochastic chemostat model with periodic washout rate, on the basis of Khasminskii’s theory (see, ref.22 Chapter 3) for periodic Markov processes. In this paper, the dynamical behaviour of a two-dimensional chemostat model with Monod-Haldane response function under stochastic perturbation is investigated. Te white noise is incorporated in stochastic system (1.3) to model the efect of a randomly fuctuating environment. Te substrate and microorganism population are usu- ally estimated by an average value plus errors (i.e. noise intensities) which are usually normally distributed. Te standard deviations of the errors may depend on the population sizes. Utilizing the approach used in17 to include stochastic efects (readers can also refer to23, Appendix A to see the construction of this kind of stochastic model), the stochastic chemostat model with Monod-Haldane response function takes the form:

    0 1 mS dS = QS()−−S xd tS+ σ11dB ()t ,   δ aS++KS2     mS  dx =  xQ− xd tx+ σ dB ()t ,   2  22  aS++KS  (1.5)

where B1(t), B2(t) are independent standard (i.e. with variance t) Brownian motion defned in a complete proba- 2 2 bility space (,Ω ,{ tt},≥0 P), and σσ1 ,02 > are their intensities. Since an ODE model is never exact, but only   an approximation of reality; adding a linear perturbation using Brownian motion (terms σ11SB , σ22xB ) as in model (1.5) helps to determine how the dynamics can possibly change by approximating the error in the ODE model, 2 and the error is normally distributed, σσii(0> ), i = 1, 2 are constants that refect the sizes of error (stochastic efects). Very few other studies have appeared on the stochastic chemostat model with the Monod-Haldane functional response. Campillo et al.24 considered a stochastic model of the chemostat, with both non-inhibitory (Monod) and inhibitory (Haldane) growth functions, as a difusion process and used a fnite diference scheme to approx- imate the solutions of the associated Fokker-Planck equation. Tey considered the stochastic model from a ‘demographic noise’ point of view. Instead of using S and x in the stochastic terms in model (1.5), they used S and x, respectively. In this paper, our aim is to reveal how the environmental noise afects a microbial population in the chemostat with Monod-Haldane response function. First, in section 2 we prove that there is a unique positive solution for 2 model (1.5). Ten in section 3 we show that for any initial value ((Sx0),(0)) ∈ +, there is a stationary distribu- tion for system (1.5) and it is ergodic under appropriate conditions. Sufcient conditions for extinction of the microorganism are presented in section 4. Finally, to illustrate our main conclusions, examples and numerical simulations are given in section 5.

SCientifiC RepOrTs | 7:13641 | DOI:10.1038/s41598-017-13294-3 2 www.nature.com/scientificreports/

Existence and uniqueness of the positive solution Before investigating the dynamical behavior of system (1.5), the existence of a global positive solution is proved. In order to ensure the stochastic chemostat make sense, we need to show at least not only that this SDE model has 2 a unique global solution but also that the solution will remain in + whenever it starts from there. First we give the defnition of a solution of stochastic diferential equation (see25, Chapter 2 for more details). Troughout this paper, unless otherwise specifed, let (,Ω ,{ tt},≥0 P) be a complete probability space with fltration {}tt≥0 sat- isfying the usual conditions (i.e. it is right continuous and 0 contains all P-null sets). Let ll T + =∈{:xxi >≤01forall il≤ } and B()tB=…((1 tB), ,(m t)) , t ≥ 0 be an m-dimensional Brownian l motion defned on the space. Let Xt()00= X (0 ≤

dX()tf=+((Xt), td)(tgXt(),)tdBt() on tt0 ≤≤T (2.1)

with initial value X(t0) = X0. SDE (2.1) is equivalent to the following stochastic integral equation:

t t =+ +≤≤ Xt() Xf00∫∫((Xs), sd)(sfXs(),)sdBs()on ttT t00t (2.2)

Defnition 2.1 An l-valued stochastic process {(Xt)} is called a solution of equation (2.1) if it has the fol-  tt0≤≤T lowing properties:

(a) {X(t)} is continuous and t-adapted; 1 l 2 lm× (b) fX((tt), )(∈  [,tT0 ];  ) and g((Xt), tt)(∈  [,0 T];  ); (c) equation (2.2) holds for every t ∈ [,tT0 ] with probability 1.

A solution {X(t)} is said to be unique if any other solution {(Xt)} is indistinguishable from {X(t)}, that is

PX{(tX)(=≤tf)}or allt0 tT≤=1, where P denotes the probability of an event. By utilizing the methods described in25, the coefcients of the equations would be required to satisfy a local Lipschitz condition and a linear growth condition. However, the Haldane function S →=μ()S mS is non- aS++KS2 linear, coefcients of system (1.5) do not satisfy the linear growth condition. Tus in this section, using the Lyapunov analysis method26, we prove that the solution of the system (1.5) is positive and global.

2 Theorem 2.1 For given initial value ((Sx0),(0)) ∈ +, there is a unique solution ((St), xt()) of system (1.5) 2 2 defned for all t ≥ 0, and the solution remains in + with probability one, i.e. ((St), xt()) ∈ + for t ≥ 0 almost surely.

Proof: Consider the difusion process as follows

  0 ν   S  11me dQμ =   − 1 − − σσ2 dt + dB ()t ,    μ  μμ2 1  11   e  δ ae++Ke 2       meμ 1  dνσ=  −−Qd2 td+.σ Bt()   μμ2 2  22  ae++Ke 2  (2.3) Since the coefcients of system (2.3) are locally Lipschitz continuous, there is a unique local solution for sys- tem (2.3). Let S = eμ, x = eμ, Itô’s formula (given in Section 3) implies that system (1.5) has a unique local positive solution. Hence it sufces to prove that this unique local positive solution of system (1.5) is global. On the basis of the discussion above, we know that there is a unique local solution (S(t), x(t)) on t ∈ [0,)τe , where τe is the blow up time. If we can prove τe = ∞ a.s., then the solution will be global. Choose n0 ≥ 0 big  1  enough in order for S(0), x(0) to lie within the interval  , n0. For each integer n ≥ n0, defne the stopping time:  n0  1 ττne=∈inft [0,):min{(St), xt()} ≤≥or max{St(),(xt)} n . {}n

Troughout this paper, we set infφ =∞ (as usual, φ denotes the empty set). Clearly, τn is increasing as →∞ 2 n . Set τ∞ = lim τn, where τ∞ ≤ τe a.s. If we prove that τ∞ =∞ a.s., then τe =∞ and ((St), xt()) ∈ + a.s. n→∞ for all t ≥ 0. In other words, to complete the proof all we need to show is that τ∞ =∞as.. If not, there exists a pair of constants T > 0 and ε ∈ (0,1) such that

PT{}τε∞ ≤>.

Hence there is an integer n1 ≥ n0 such that

SCientifiC RepOrTs | 7:13641 | DOI:10.1038/s41598-017-13294-3 3 www.nature.com/scientificreports/

PT{}τεn ≤>, forall nn≥.1 (2.4)

2 Defne a C -function VS(,x): ++×→ by S 1 VS(,xS)l=−AA−+og (1xx−−log), A δ where A = Qa . Obviously, this function is non-negative for all S, x ≥ 0. Using Itô’s formula (see Teorem 3.1), we m get

SA−  1 mSx   1 dV(,Sx)(= QS0 −−S) dt +−σσ()SAdB ()tA + 2 dt  2  11 1 S  δ aS++KS   2 1  mSx   1 +−(1x )( − Qxdt + σ dB td) + σ 2 t  2  22 2 δ aS++KS   2δ 1 :(=+VS,)xdtSσ ()−+AdBt() σ (1xd− )(Bt), 11δ 22 where  is the generating operator of system (1.5) and

 A 0 1 mSx Amx VS(,xQ)1=  − ()SS−− +  S  δ aS++KS22aS++KS

xm− 11 S  2 1 2 +  − QA ++σ1 σ2 δ aS++KS2  2 2δ 0 Am Q Q 1 2 1 2 ≤+QS −+x ++σ1 A σ2 aδδ δ 2 2δ 0 QQa 2 1 2 =+QS ++σ1 σ2 δ 2m 2δ  C⁎, C* is a positive constant. Terefore

ττ∧∧TTτ ∧T nn⁎ n dV((St), xt())(≤+CdtSσ11− Ad)(Bt) ∫∫00∫0 τ ∧ n T 1 +−σ22(1xd)(Bt), ∫0 δ Taking expectation of both sides implies that

τ ∧T n ⁎ EV((STττnn∧∧), xT())(≤+VS(0), xE(0)) Cdt ∫0 ≤+VS((0),(xC0)) ⁎T. (2.5)

Set Ωnn=≤τ T for all n ≤ n1. Ten by (2.4) P()Ω≥n ε. Note that for every ω ∈Ωn, there exists at least one S(,τω) or x(,τω) that equals either n or 1 , and then VS((ττ), x()) is no less than n n n nn 1 11 nn−−1log or −−1log =−1l+.ogn nnn Consequently,   1  VS((ττnn), xn())(≥−1l−∧ogn)  −+1logn. n  It then follows from (2.4) and (2.5) that

VS((0),(xC0))(+≥⁎TE1 VS()ττ,(x ))  Ωωn() nn  1  ≥ ε(1nn−−log)∧  −+1logn,  n 

where 1Ω ()ω is the indicator function of Ωn. Letting n → ∞ leads to the contradiction that ∞ > VS((0),(x 0)) ⁎ n +CT=∞. Tus we have τ∞ = ∞ a.s. ☐ Stationary distribution and ergodicity In the study of the dynamics for stochastic systems, ergodicity is one of the most important and signifcant char- acteristics. Te ergodic property for chemostat implies that the stochastic model has a unique stationary distribu- tion which predicts the survival of the microbial population in the future. In addition, the ergodic property gives

SCientifiC RepOrTs | 7:13641 | DOI:10.1038/s41598-017-13294-3 4 www.nature.com/scientificreports/

a reason why the integral average of a population system converges to a fxed point whilst the population system may fuctuate around as time goes by. Tus research on the ergodicity of population system is essential from a bio- logical perspective (see27). In this section, before the proof of the ergodicity, several auxiliary results are derived for the stationary distribution. For more details, we refer the reader to28. First we introduce the multi-dimensional Itô formula.

Teorem 3.1 (Te multi-dimensional Itô formula) (Chapter 1)25 Let X(t) be a l-dimensional It oˆ process on t ≥ 0 with the stochastic diferential dX()tf=+()tdtg()tdBt(),

1 d 2 dm× 2,1 d where f ∈  (;+ ) and g ∈  (;+ ). Let VC∈×(;+ ). Ten V(X(t), t) is an Itoˆ process with the stochastic diferential given by

1 T dV((Xt), tV)[=+tX((Xt), tV)(Xt(),)tf()tt+ race ((gt)(VXXX ()tt,)gt( ))]dt 2 +.VXX((tt), )(gt)(dB ta) s. (3.1)

∂V ∂V ∂V  ∂2  Here V = , V =…,, , V =  V  Assume Y(t) is a regular time-homogeneous Markov t ∂t X ∂X ∂X XX  ∂∂XX ()1 l  ij × l l ll Process in  ( denotes the l-dimensional Euclidean space) described by the stochastic diferential equation

l =+ dY()tb()Ydtg∑ r()YdBtr(), r=1 (3.2)

Brownian motion Br, r = 1, …, l are independent. Te difusion matrix is

l Λ=λλ=.i j ()y ((ij yyg)),(ij )(∑ r yg)(r y) r=1 Defne the diferential operator  associated with equation (3.2) by

l ∂ 1 l ∂2  = ∑∑byk() + λkj()y . ∂ ∂∂ k==1,yk 2 kj 1 yykj

l Lemma 3.1 (Rafail Khasminskii, 2011) Assume that there exists a bounded domain U ⊂  with regular bound- ary Γ, having the following properties: (B.1) In the domain U and some neighborhood thereof, the smallest eigenvalue of the difusion matrix Λ()x is bounded away from zero. ∈ l τ (B.2) If x  \U, the mean time for the paths issuing from x to reach the set U is fnite, and sup Exτ <∞ for ∈ l xD every compact subset D ⊂  . Here E means expectation. Ten the Markov process X(t) has a stationary distribution μ()⋅ . Moreover, let f(x) be a function integrable with respect to the measure μ. Ten the probability 1 T P{lim fX((td)) tf==()xdμ()x }1, T→∞T ∫∫0 El l for all x ∈  .

Remark 3.1 Te proof is given in22. To show the existence of the stationary distribution μ()⋅ of system (1.5), it is 2 enough for us to take + as the whole space. To validate (B.1), we need to prove that for any bounded domain ⊂ 2 2 λξξξ≥||∈22ξ ∈ 29 D +, there is a positive number M0such that ∑ij,1= ij()xMij 0 ,,xD R+ (see Chapter 3, p.103 and Rayleigh’s principle in30 Chapter 6, p.349). To validate (B.2), it is sufcient to show that there exists a neighbor- l hood U and a non-negative C2-function V(S, x) such that for any  \U, VS(,x) is negative (for more details see31, p.1163).

Defnition 3.1 (Regularity, Recurrence) A Markov process X(t) is said to be regular, if for any 0 <

P{}τx <∞ = 1,

where τx =>inf{tX0: (0),=∈xX()tU} is the hitting time of U for X(t).

SCientifiC RepOrTs | 7:13641 | DOI:10.1038/s41598-017-13294-3 5 www.nature.com/scientificreports/

Remark 3.2 Te defnition of regularity and recurrence come from32. Let X(t) be a regular temporally homogeneous l Markov process in  . If X(t) is recurrent with respect to some bounded domain U, then it is recurrent with respect to l any nonempty domain in  . Since the existence of the positive solution for model (1.5) has been obtained by Teorem 2 2.1, it is enough for us to take + as the whole space. Before proving the ergodicity, we defne the notation ffu = sup(t), ffl = inf(t). Defne a continuous t∈∞[0,) t∈∞[0,) function G: + → by

~~02 ~ 00 GS() =−()QQ−−σσ1 KS()SQ+−1(2mQ+−KS )(SS) 00~ 02 −−σ1[(mS Qa++SKS )]. where Q =+Q 1 σ 2. 2 2

Teorem 3.2 If σ1, σ2satisfy:

2 −2C σσ10<

~ mS0 < σ 2 < Q 2 ,22 Q, aS++00KS (3.4) ~ GS() Q 4[QmSQ00−+()aS+ KS02 ] where C =

Proof: For simplicity let S, x stand for S(t), x(t), respectively. We will verify that (B.1) and (B.2) hold under condi- tion (3.3–3.4), according to Lemma 3.1 on the existence of the stationary distribution. System (1.5) can be written as the following system:

 0 1 mS  QS()−−S x  δ ++ 2    S  aSKS  σ1S d = dt + dB1()t ()x  mS   0   xQ− x   aS++KS2  0 + σ dB2()t , ()2x hence the difusion matrix is

 22  σ S 0  Λ=(,Sx)  1 .  22 0 σ2 x 

2 22 Let D to be any bounded domain in +, then there exists a positive constant M01= min{σ S , 22 σ2 xS,( ,)xD∈ } such that 2 λξξσ=+22ξσ2 22ξξ2 ≥||∈22ξ ∈. ∑ ij(,Sx),ij 1 Sx1 2 2 Mf0 or allS(,xD), R ij,1= Tis implies that condition (B.1) is satisfed. 2 2 Next we will construct a nonnegative C -function V(S, x) and a closed set U ∈ + such that sup(VS,)xR<− < 0, 2 (,Sx)\∈+ U where R is a positive constant. Tis assures that condition (B.2) is satisfed. Choose a positive constant M big enough such that 0 (A) Φ−u Mλ ≤−2,where λ =−C −>S σ 2 0 and the function Φ()S will be given later in (B). Defne a 2 1 C2-function:

   2 00S 1  0 1  HS(,xM)l= SS−−S og − σ1 logx + SS−+xS −.log  S0  2Q  δ  From the partial derivative equation of H(S, x) we have

0 11−+MS MS  0 MSσ  − + SS−+ 1  =.0 δδSQ 1 −+MS MS0  (3.5)

SCientifiC RepOrTs | 7:13641 | DOI:10.1038/s41598-017-13294-3 6 www.nature.com/scientificreports/

We derive the unique solution S0 of (3.5) from the monotonicity property of the lef part. Tus the following equation x S = 0 Mσδ1 1 −+MS MS

MSσδ10 has a unique solution (S0, x0) which is the minimum point of function H(S, x). Here x0 = 0 . Terefore 1 −+MS0 MS H(,Sx)(−≥HS00,)x 0. 2 2 Next, we defne a nonnegative C -function V(,⋅⋅): ++→ by    2 00S 1  0 1  VS(,xM)l= SS−−S og − σ1 logx + SS−+xS −−log(HS00,)x .  S0  2Q  δ  2 00S 1 0 1 Denote VS(,xS)l=−SS− og 0 , VS21(,xV)(=−Sx,) σ1 logx, VS(,xS) =−Sx+ . 1 S 3 2Q ()δ Hence

VS(,xM)(=+VV23+−log)S . (3.6) By Itô’s formula, we obtain that

0 0 SS−  0 1 mSx  S 2 VS1(,x)(= QS −−S)  + σ1 S  δ aS++KS2  2 02 0 0 QS()− S 1 mS x S 2 ≤− + + σ1 S δ aS++KS2 2 02 00 QS()− S mS S 2 ≤− ++x σ1 S aδ 2 02 00 QS()− S mS S 2 ≤− ++x σ1 , aS++KS2 aδ 2 and ~ mS (l−=og xQ) − aS++KS2 ~~~ Qa ++QS QKSm2 − S = aS++KS2 ~~~~2 QK()SS−+02 (2Qm−+QKSS00)( −−Sm)[SQ−+()aS00+ KS ] = . aS++KS2 Tus

~~~~2 σσQK()SS−+02 (2Qm−+QKSS00)( −−Sm)[σ SQ−+()aS00+ KS ] VS(,x) ≤ 1 1 1 2 aS++KS2 02 00 QS()− S mS S 2 − ++x σ1 aS++KS2 aδ 2 ~~~~2 −−σσ−+02 −+ 00−−σ −+00+ = ()QQ1 KS()SQ1(2mQKS )(SS)[1 mS Qa()SKS ] aS++KS2 00 mS S 2 ++x σ1 aδ 2 00 GS() mS S 2 = ++x σ1 . aS++KS2 aδ 2  It follows from the condition (3.3) that the discriminant of function G(S) is negative and Q −>σ1QK 0. Ten 0 GS() S 2 for any S ∈∞(0,), function G(S) < 0. Take C =σ1 0, then we get aS++KS2 2 S∈∞(, ) mS0 VS2(,x) ≤−λ +.x aδ (3.7) 2 In addition, since S20≤−2(SS)220+ S , we have

SCientifiC RepOrTs | 7:13641 | DOI:10.1038/s41598-017-13294-3 7 www.nature.com/scientificreports/

2 2 021 σσ1 2 2 2 VS3(,xS)(=− −+Sx) ++S x δ 2Q 2Qδ2 12 σσ2 2 =−()SS−−02 xS20−−()Sx++1 S2 2 x2 δ2 δ 2Q 2Qδ2 2 2 002 σ1 02 1 σ2 2 2S S 2 ≤−(1 −−)(SS) −−(1 )x ++x σ1 , Q δ2 2Q δ Q (3.8) and

0 QS 11mx 2 (l−=ogS) −+Q + + σ1 S δ aS++KS2 2 0 QS m 1 2 ≤− ++Q x +.σ1 S aδ 2 (3.9) Substituting (3.7–3.9) into (3.6),

mS0 σ 2 1 σ 2 VS(,xM)(≤−λ +−x)(1)−−1 ()SS02−−(1 2 )x2 aδ Q δ2 2Q 00 02 21S QS m 2 S 2 +−x ++Q x ++σσ1 1 δδS a 2 Q =Φ()Sx+Ψ(), where

 2 0 02  σ1 02 QS 1 2 S 2 Φ=()S −−(1 )(SS−−) ++Q σσ1 + 1  Q S 2 Q ()B   1 σ 2 2S00m mS Ψ=−−2 2 +++−λ +.  ()x 2 (1 )x x xM()x  δ 2Q δδa aδ (3.10) In view of (3.3) and (3.4), we observe that

Φ+()SSΨ→u −∞,as,→+∞ Φ+u Ψ→()xx−∞,as,→+∞ Φ+()SSΨ→u −∞,as0→.+ Te above cases lead to V <−1, respectively. Moreover, by reviewing condition (A) we obtain

Φ+uuΨ→()xMΦ− λ ≤−2, as x →.0+

Take ε small enough, and let U = εε,,11 ×  . Terefore  εε   VS(,xS)1<− ,(,)xU∈.c According to Remark 3.1, the conditions in Lemma 3.1 are satisfed. Tus stochastic chemostat model (1.5) has a stationary distribution and it is ergodic. ☐ Extinction In this section, we discuss conditions that predict the failure of the continuous culture of microorganisms both in the case of environmental noise is ignored and in the case of big intensities of white noises, i.e. noise may lead to extinction of the microorganism in the reactor.

Teorem 4.1 Let Q =+Q 1 σ 2. If one of the following conditions holds 2 2 (i) Q ≥ m, 2 (ii) Q < m, and (1 −−4)aK QQ20mm+<2 . 2 Ten for any given initial value ((Sx0),(0)) ∈ +, the solution ((St), xt()) of system (1.5) satisfes 1 limsup log(xt())0<.as., t→∞ t (4.1) which means x(t) tends to zero exponentially almost surely. In other words, the microorganisms die out with proba- bility one.

SCientifiC RepOrTs | 7:13641 | DOI:10.1038/s41598-017-13294-3 8 www.nature.com/scientificreports/

Proof: By Itô’s formula, we have

t  mS 1 2 log(xt)l=+ogx0  −−Qdσσ tB+ 22()t , ∫0 aS++KS2 2  and this implies that

1 1(t  mS r) 1  limsup log(xt())l= imsup  −−Qdσ2 ra..s . ∫  2  tt→∞ t →∞ t 0  aS++()rKSr() 2  −−2 −− Denote g()SQ=−mS . If (i) holds, then g()SQ=

Remark 4.1 We refer to the condition (ii) in Teorem 4.1, which tells us that the microorganism species may die out when dilution rate Q and white noise are not large. While if the strength of white noise is large enough such that (i) holds, then the microorganism population will also become extinct, which never happens in the deter- ministic system (1.3) without environmental perturbations. Moreover, if the dilution rate Q is big enough which leads to the extinction of the species in the deterministic chemostat while the noise is not big (condition (i)), the microorganism species in stochastic chemostat (1.5) will also die out. From Teorem 4.1, under either condition (i) or (ii), we obtain that there is some constant λ > 0 such that 1 limsup log(xt()) ≤−λ. t→∞ t λ Tat is to say, for a arbitrary small constant 0m<<ε in 1 , , there exists a positive constant T = T()ω and {}22 11 λ −λt a set Ω such that P{}Ω≥1 − ε and logxt()≤− for t ≥ T , ω ∈Ω. Ten x()te≤ 2 . Tus ε ε 2 1 ε limsupxt()≤.0, as., t→∞ which together with the positive property of the solution of system (1.5) implies that lim(xt)0=., as. t→∞ Terefore the microorganism species x will go to extinction exponentially almost surely. In other words, the micro- organism population will die out at an exponential rate with probability one. Examples and numerical simulations Generally, nonlinear stochastic diferential equations, such as system (1.5), are usually too complex to be solved exactly, we can only theoretically prove the existence and uniqueness of the positive solution (see Teorem 2.1). In order to illustrate the analytical results in Section 3 and 4, we need to obtain approximate solutions of stochas- tic system (1.5) with given initial values and parameter values, via numerical methods and algorithms that can be implemented by Matlab. Due to Brownian motion and noise terms σ11SdBt(), σ22xdBt(), we use Milstein’s higher order method33 to obtain the approximate solutions of stochastic chemostat model (1.5). J. Higham has showed the convergence of Milstein’s numerical method for stochastic diferential equations (see Section 6, ref.33). We consider the following discretization equation in Milstein’s type to iteratively calculate the approximate solutions of stochastic system (1.5) in Matlab programs:    1 mS x   =+ 0 −− ii ∆ SSii+1 QS()Si 2  t   δ aS++iiKS    σ 2  +∆σξ +∆1 ξ 2 −∆  1Sti 1,i Sti()1,i t ,  2   mS x  xx=+ ii − Qx ∆t  ii+1  2 i  +aSii+ KS    σ 2  +∆σξ +∆2 ξ 2 −∆  2xti 2,i xti()2,i t ,  2 (5.1)

SCientifiC RepOrTs | 7:13641 | DOI:10.1038/s41598-017-13294-3 9 www.nature.com/scientificreports/

Figure 1. Numerical simulations of the solutions for system (1.5) and the corresponding deterministic system (1.3) with main parameter Q = 1. Both S(t) and x(t) are stationary Markov processes. Te subgraphs on the right are the density functions of the correspondng stationary distributions for low level intensity Brownian motion with σσ12=.005, =.01.

ξ ξ σσ where 1,i and 2,i are N(0,1)-distributed independent Gaussian random variables, 12, are intensities of white noise and time increment ∆>t 0. In this way, the approximate solutions (performance results of (5.1) from Matlab) will converge to the explicit solutions of stochastic system (1.5) very fast.

Example 5.1 From Teorem 3.2, we expect that under some appropriate conditions, the stochastic system (1.5) has a stationary distribution. Choose the initial value ((Sx0),(0))(=.12,0.8) and assume that the parameters in 0 (1.5) are given by Q = 1, m = 2, S = 2, a =.02, K =.03, δ =.16, and choose σ1 =.005, σ2 =.01. Hence conditions:

2 2C 00.=53σσ10<=G ..1192,00025 =<1 −∧Q =.0 0086, S0

~ mS0 .=< =. .=σ 2 <= 1 0050 Q 2 1 1765,0 01 2 22Q , aS++00KS in Teorem 3.2 are all satisfed. Ten we conclude that there is a stationary distribution (see the right two sub- graphs in Fig. 1) of the stochastic system (1.5) and it is ergodic. Numerical simulations in Fig. 1 support this conclusion clearly, illustrating that the standard deviation σ keeps processes S(t), x(t) moving around the solution of the corresponding deterministic chemostat model.

Example 5.2 To further illustrate the efect of the Brownian motion σ on the stochastic chemostat model, we keep all the parameters in Example 5.1 unchanged but increase intensities to σ1 =.01, and σ2 =.02. We compute

2 2C 01.=σσ10<=G 2..9942, 001 =<1 −∧Q =.0 0156, S0

SCientifiC RepOrTs | 7:13641 | DOI:10.1038/s41598-017-13294-3 10 www.nature.com/scientificreports/

Figure 2. Numerical simulations of the solutions for system (1.5) and the corresponding deterministic system (1.3). Te intensity of the Brownian motion is increased to σσ12=.01,0=.2. Te stochastic chemostat model still has a stationary distribution, even though the amplitude oscillations are stronger.

~ mS0 .=< =. .=σ 2 <= 1 0200 Q 2 1 1765,004 2 22Q , aS++00KS which means these parameter values satisfy the ergodic conditions (3.3–3.4). Figure 2 illustrates that the white noise still keeps processes S()t , x()t moving around the solution of the deterministic chemostat model, while the random oscillations become stronger.

Example 5.3 In order to show how the dilution rate and the random perturbation infuence the extinction of the microbial population in system (1.5), we choose the initial value ((Sx0),(0))(=.3, 05) and assume that Q =.22, σ1 =.03, σ2 =.04 with the other parameters unchanged. Tus

1 2  2.=280 QQ+ σ2 =>=m 2. 2 We therefore conclude that by Teorem 4.1 (i), the solution of (1.5) obeys 1 limsup log(xt())0<−Qa <.s., t→∞ t that is, x(t) tends to zero exponentially with probability one. On the other hand, for the corresponding deter- ministic chemostat model (1.3), condition

mS0 .= > =. 22 Q 2 1 1765 aS++00KS is satisfed, so the trivial equilibrium point (,S0 0) = (2,0) is a stable nodal point. Terefore for these parameters, x()t in system (1.3) also dies out. Numerical simulation in Fig. 3 provides clear support.

SCientifiC RepOrTs | 7:13641 | DOI:10.1038/s41598-017-13294-3 11 www.nature.com/scientificreports/

Figure 3. Te sample paths of S(t) and x(t) for system (1.5) and the corresponding deterministic system (1.3) with Q = 2.2. Te intensity of Brownian motion is chosen to be σσ12=.03,0=.4. In the second subgraph, the large washout rate leads to the extinction of x(t) both in the deterministic and stochastic chemostat models.

Figure 4. Te sample paths of S(t) and x(t) for system (1.5) and the corresponding deterministic system (1.3) with Q = 1.6. For the intensity of the Brownian motion chosen to be σσ12=.03,0=.4. In this case, condition (ii) in Teorem 4.1 is satisfed. Te microorganism population dies out in both cases.

Example 5.4 We decrease the dilution rate Q to 1.6, and choose the same noises densities as that in Example 5.3, then

1 2 2  2 1.=680 QQ+ σ2 =<=ma2, (1 −−4)KQ 20Qm +=m −.5750 <.0 2 Terefore, by Teorem 4.1 (ii), the solution of (1.5) obeys 1 limsup log(xt())1<− .<6798 0,as.. t→∞ t and so x()t tends to zero exponentially with probability one. Tese results are illustrated in the numerical simula- tions in Fig. 4. We can observe from Figs 3–4 that under the same random noises, the extinction time of the microbial population occurs later when Q =.16 than for Q =.22 for both the deterministic system (1.3) and the stochastic system (1.5).

Example 5.5 As a comparison, we reduce the volumetric fow rate Q to 1.1, and increase the intensity of Brownian motion B2()t to σ2 =.15 with other parameters and initial value unchanged. Condition

SCientifiC RepOrTs | 7:13641 | DOI:10.1038/s41598-017-13294-3 12 www.nature.com/scientificreports/

Figure 5. Te sample paths of S(t) and x(t) for system (1.5) and its corresponding deterministic system (1.3) with Q = 1.1, while the intensities of Brownian motion are increased to σσ12=.03,1=.5. Te large scale of white noises cause the death for x()t in the stochastic chemostat model, while the deterministic one is survival.

Q ↓ (1), Q ↑ (2.2), Microbial population σ2 ↓ (0.1) Q ↓ (1.1), σ2 ↑ (1.5) Q ↑ (1.6), σ2 ↓ (0.4) σ2 ↓ (0.4) Deterministic system Survival Survival Extinction (slowest) Extinction Stochastic system Survival Extinction (fastest) Extinction (slowest) Extinction

Table 1. Efects of the dilution rate and stochasticity on the survival-extinction of the microorganism.

1 2  2.=2250 QQ+=σ2 >=m 2 2 of Teorem 4.1 (i) is satisfed, so the solution of system (1.5) obeys 1 limsup log(xt())0<−Qa <.s., t→∞ t which means x()t of system (1.5) will go to extinction with probability one. Note that

mS0 .= < =. . 11 Q 2 1 1765 aS++00KS In this case, the corresponding deterministic chemostat model (1.3) has only one interior positive equilibrium E⁎ =.(0 2715,2.7657) and E* is a globally asymptotically stable node. Te numercial simulations in Fig. 5 con- frm the extinction of x()t in the stochastic system (1.5) and the stability of E* in the deterministic system (1.3), and it is clear that with the increasing of the intensity of white noise, the tendency for exponential extinction increases. Discussion and Conclusion Understanding the efects of environmental stochasticity on the survival of microbial populations is of theoretical and practical importance in population biology. In this paper we consider a stochastic chemostat model with the Monod-Haldane response function. We frst determine when this system has a unique global positive solution. Ten we show that the stochastic chemostat model admits a unique stationary distribution which is ergodic if the scale of the random perturbations is relatively small. Sufcient criteria are provided for the extinction of the population of microorganisms. We also fnd that both strong random noise and a large dilution rate can lead to extinction. Our conclusions are all expressed in terms of the system parameters and the intensity of the Brownian motion. Tis means that white noise can have a major impact on the survival or extinction of microorganism populations. How does environmental stochasticity change the predicted outcome for the microbial population in (1.3)? Actually, the stochastic chemostat (1.5) has no equilibrium. Te stationary distribution shows that the solutions of the stochastic system fuctuate in a neighborhood of the positive equilibrium of the corresponding determin- istic system, which can be regarded as weak stability. In Teorem 3.2, we defne a new dilution rate Q which is in terms of the original dilution rate Q and the random noise on the microbial population σ , i.e. Q =+Q 1 σ 2. For 2 2 2 the parameter conditions in16 and some restrictions on the intensity of the white noise, the stochastic system (1.3)

SCientifiC RepOrTs | 7:13641 | DOI:10.1038/s41598-017-13294-3 13 www.nature.com/scientificreports/

Figure 6. Vector feld and trjactories of deterministic chemostat model (1.3) and the dynamical behavior of the stochastic chemostat model (1.5) with Q = 0.7. In this case, the parametric values do not satisfy the conditions in Teorem 3.2. In subgraph (a), both the boundary equilibrium E0 and the positive equilibrium E1 are locally 2 stable nodes while E is a saddle point. Choosing the intensity of Brownian motion to be σσ12=.005, =.005 relatively small result in small fuctuations. Tese solution trajectories in (b) have similar properties to that of the deterministic model.

has a unique stationary distribution and it is ergodic, which means the microbial population survives for all time regardless of the initial conditions. Tis result illustrates that random noise in low levels is advantageous for spe- 2 cies survival. In Figs 1 and 2, numerical simulations show the dynamics of model (1.5) with diferent σi , i = 1,2. 2 Comparing Fig. 1 with Fig. 2 we see that with increasing σi , the oscillations become stronger, but both fgures reveal that the microbial population persists and demonstrate the existence of the unique stationary distribution. With increasing Q, Teorem 4.1 shows that x()t tends to extinction exponentially. Furthermore, it can be observed from Figs 3–5 that the extinction of the microbial population occurs more quickly for system (1.5) with σ2 =.15 than that with σ2 =.04. However, with respect to the corresponding deterministic model, x()t in system (1.3) with Q =.22 tends to extinction faster than that with Q =.16. Tat is, if the volumetric fow rate Q is increased, the microbial population tends to extinction faster. Thus from the biological point of view, a well-controlled volumetric fow rate in the chemostat is also the key to getting a successful continuous culture of microorganisms. We have already shown that if the densities of the environmental noise are small, the deterministic and sto- chastic systems have similar dynamical behavior, both in the case of persistence and extinction. Taking the dilu- tion rate Q =.11 in system (1.3) with σi = 0, i = 1, 2, there exists an asymptotically stable interior equilibrium. Increasing σ2, the stochastic trajectory x()t will frst fuctuate around this equilibrium, then the phenomenon of noise-induced extinction occurs when σ2 is big enough. In this stochastic case, the microorganism tends to extinction exponentially even faster than in the case of large dilution rate Q =.22. Tese results are summarized in the following Table 1. To complement our theoretical result we study the global dynamics of the deterministic chemostat system (1.3) and the stochastic chemostat system (1.5) in cases that are not counter by our theorems. We set parametric 0 =. 0 =. ==δ =. =. mS values Q 07,2Sm4, 3, 1, aK09,16. In this case, 07.=Q >=2 0.5753. Ten aS++00KS according to the mathematical results in16, the deterministic chemostat model (3) has two interior equilibrium points, one an asymptotically stable node E1 =.(0 3255,2.0745) and one a saddle point E2 =.(1 7281,0.6719), as well as one boundary equilibrium E0 =.(2 4, 0) that is also an asymptotically stable node. Te topological struc- ture of the deterministic chemostat model (1.3) in Fig. 6(a) shows that E0 and E1 are both stable. Te microbial population will survive if and only if the orbits tend to the interior stable node E1. Tis means in order for the population to survive, we must choose appropriate initial concentrations of nutrient and microbial population. For the corresponding stochastic chemostat model (1.5), we set σ1 =.005, σ2 =.005. With the same initial con- ditions and parameter values, Fig. 6(b) suggests that processes S()t , x()t move around the trajectories of the deter- ministic chemostat model (1.3) with small fuctuations. Tey have similar properties to that of the deterministic model. Note that the existence conditions for a stationary distribution are only sufcient, but not necessary. In case of Q =.07, these parameters do not satisfy the hypotheses of Teorem 3.2, and so we cannot estimate whether the stochastic chemostat is ergodic or not. Tis is an open problem.

SCientifiC RepOrTs | 7:13641 | DOI:10.1038/s41598-017-13294-3 14 www.nature.com/scientificreports/

For biological population models, besides white noise, there is another type of environmental noise called color noise, or telegraph noise. Color noise can be illustrated as a switching between two or more environmental regimes which difer by factors such as humidity and temperature (see e.g.32,34). For example, the growth rates of some species in the dry season will be much diferent from those in the rainy season. Color noise is utilized to describe this phenomenon of environmental regimes. Color noise will disturb the steady-state of a real system indirectly by afecting system parameters such as growth rate and death rate. Tis is diferent from the perturba- tions of white noise that directly act on population densities (see system (1.5)). From the observation point of view, the random efect of Brownian motion is more visualized (for example, the irregular movement of pollen grains). For color noise, the switching is memoryless and the waiting time for the next switch is exponentially distributed. Hence the regime switching can be modelled by a continuous-time Markov chain rt(), t ≥ 0 with fnite-state space  = {1,2,, n}. In view of the feasibility of varying the dilution rate in a chemostat (see5), it is also realistic to introduce color noise, or regime switching, into Q. Ten the value of dilution rate Q((rt)) will switch according to the law of Markov chain rt(). For bio-mathematical model, many researchers have studied the dynamics under both white and color noise (see27 and the references there in). If color noise is taken into account, then we should study the ergodic property of the stochastic chemostat with Monod-Haldane growth function and πσ < π regime switching. The sufficient conditions for ergodicity are supposed to be ∑k∈∈ k 10()kG∑k  k , 2C 0 ∑ πσ2()kQ<−∑ ππ∧ ∑ ()k ; and ∑ πσ((Qk)(+<1 2 k)) mS , k∈∈ k 1 k  k 0 k∈ k k∈ k 2 002 S 2 aS++KS πσ2 < π ∑k∈∈ k 2 ()kQ∑k  k ()k , which are all expressed in terms of system parameters, the intensities of Brownian motion and the distribution for Markov chain, i.e. the mathematical conclusions in Teorem 3.2 should be the mean value on space average. While in the proof of Teorem 3.2, the key is to make sure that λ is positive, accord- ing to condition (3.3). For stochastic system (1.5) with switching regime, it would be too difcult to fnd the mean value of λ due to the inhibitory efect. Terefore in this paper, we only introduce white noise into deterministic chemostat model (1.3). Some other interesting topics deserve further investigation. Motivated by the work in5,21, one may extend our results to periodic solutions for the periodically forced stochastic chemostat model. One may concentrate on simple microbial interactions such as competitive coexistence and predation with stochasticity. We can also take other kinds of environmental noise into account, for example Lévy noise (see e.g.35). We leave these questions for future researches. References 1. Smith, H. L. & Waltman, P. The Theory of the Chemostat: Dynamics of Microbial Competition. Cambridge University Press, Cambridge (1995). 2. Wolkowicz, G. S. K. & Zhao, X. N-Species Competition in a Periodic Chemostat. Difer. Integral. Equa. 11, 465–491 (1998). 3. Monod, J. La technique de la culture continue, théorie et applications. Annales de I’Institut Pasteur 79, 390–410 (1950). 4. Novick, A. & Szilard, L. Description of the chemostat. Science 112, 715–716 (1950). 5. Butler, G. J., Hsu, S. B. & Waltman, P. E. A mathematical model of the chemostat with perodic washout rate. SIAM J. Appl. Math. 45, 435–449 (1985). 6. Xia, H., Wolkowicz, G. S. K. & Wang, L. Transient oscillation induced by delayed growth response in the chemostat. J. Math. Biol. 51, 489–530 (2005). 7. Stephanopoulos, G. & Fredrickson, A. Extinction probabilites in microbial predation: A birth and death approach. Bull. Math. Biol. 43, 165–181 (1981). 8. Smith, H. L. Competitive coexistence in an oscillating chemostat. SIAM J. Appl. Math. 40, 498–522 (1981). 9. Lenas, P., Tomopoulos, N., Vayenas, D. & Pavlou, S. Oscillations of two competing microbial populations in confgurations of two interconnected chemostats. Math. Biosci. 148, 43–63 (1998). 10. Weedermann, M., Wolkowicz, G. S. K. & Sasara, J. Optimal biogas production in a model for anaerobic digestion. Nonlinear Dynam. 81, 1097–1112 (2015). 11. Meng, X., Wang, L. & Zhang, T. Global dynamics analysis of a nonlinear impulsive stochastic chemostat system in a polluted environment. J. Appl. Anal. Comput. 6, 865–875 (2016). 12. Andrews, J. A mathematical model for the continuous culture of macroorganisms untilizing inhibitory substrates. Biotechnol. Bioeng. 10, 707–723 (1968). 13. Wang, F., Hao, C. & Chen, L. Bifurcation and chaos in a Monod-Haldane type food chain chemostat with pulsed input and washout. Chaos & Soliton Fract. 32, 181–194 (2007). 14. Pang, G., Wang, F. & Chen, L. Analysis of a Monod-Haldane type food chain chemostat with periodically varying substrate. Chaos & Soliton Fract. 38, 731–742 (2008). 15. Baek, H., Younghae, D. & Saito, Y. Analysis of an impulsive predator-prey system with Monod-Haldane functional response and seasonal efects. Math. Probl. Eng. 2009, 543187, https://doi.org/10.1155/2009/543187 (2009). 16. Chen, L. & Chen, J. Nonlinear Biodynamical Systems (Chinese copies). Science Press, Beijing (1993). 17. Imhof, L. & Walcher, S. Exclusion and persistence in deterministic and stochastic chemostat models. J. Difer. Equ. 217, 26–53 (2005). 18. Campillo, F., Joannides, M. & Larramendy-Valverde, I. Stochastic modeling of the chemostat. Ecol. Model. 222, 2676–2689 (2011). 19. Xu, C. & Yuan, S. An analogue of break-even concentration in a simple stochastic chemostat model. Appl. Math. Lett. 48, 62–68 (2015). 20. Zhao, D. & Yuan, S. Critical result on the break-even concentration in a single-species stochastic chemostat model. J. Math. Anal. Appl. 434, 1336–1345 (2016). 21. Wang, L., Jiang, D. & O’Regan, D. Te periodic solutions of a stochastic chemostat model with periodic washout rate. Commun. Nonlinear Sci. Numer. Simulat. 37, 1–13 (2016). 22. Khasminskii, R. Stochastic stability of diferential equations (second edition). Springer, New York (2012). 23. Xu, C. & Yuan, S. Competition in the chemostat: A stochastic multi-species model and its asymptotic behavior. Math. Biosci. 280, 1–9 (2016). 24. Campillo, F., Joannides, M. & Larramendy-Valverde, I. Approximation of the Fokker-Planck equation of the stochastic chemostat. Math. Comput. Simulat. 99, 37–53 (2014). 25. Mao, X. Stochastic Diferential Equations and Applications. Horwood Publishing, Chichester, UK (2008). 26. Gray, A. et al. A stochastic diferential equation SIS epidemic model. SIAM J. Appl. Math. 71, 876–902 (2011).

SCientifiC RepOrTs | 7:13641 | DOI:10.1038/s41598-017-13294-3 15 www.nature.com/scientificreports/

27. Liu, H., Li, X. & Yang, Q. Te ergodic property and positive recurrence of a multi-group Lotka-Volterra mutualistic system with regime switching. Syst. Control Lett. 62, 805–810 (2013). 28. Ji, C., Jiang, D. & Shi, N. Dynamics of a multigroup SIR epidemic model with stochastic perturbation. Automatica 48, 121–131 (2012). 29. Gard, T. C. Introduction to Stochastic Diferential Equations. Marcel Dekker INC, New York (1988). 30. Strang, G. Linear Algebra and Its Applications. Tomson Learning INC, London (1988). 31. Zhu, C. & Yin, G. Asymptotic properties of hybrid difusion systems. SIAM J. Control. Optim. 46, 1155–1179 (2007). 32. Zhu, C. & Yin, G. On competitive Lotka-Volterra model in random environments. J. Math. Anal. Appl. 357, 154–170 (2009). 33. Higham, D. J. An algorithmic introduction to numerical simulation of stochastic diferential equations. SIAM Rev. 43, 525–546 (2001). 34. Slatkin, M. Te dynamics of a population in a Markovian environment. Ecology. 59, 249–256 (1978). 35. Bao, J. & Yuan, C. Stochastic population dynamics driven by Lévy noise. J. Math. Anal. Appl. 391, 363–375 (2012). Acknowledgements The work was supported by NSFC of China (No: 11371085), and the Fundamental Research Funds for the Central Universities (No: 15CX08011A), and the Scholarship Foundation of China Scholarship Council (No: 201606620033). Te research of Gail S.K. Wolkowicz was partially supported by Natural Sciences and Engineering Council (NSERC) Discovery and Accelerator supplement grants of Canada. Author Contributions Prof. Daqing Jiang developed the study concept and design. Liang Wang analyzed and drafed the manuscript. Prof. Wolkowicz improved the written English and made many suggestions on the figures. Prof. O’Regan corrected some spelling mistakes and inaccuracy of expressions. All authors read and approved the final manuscript. Additional Information Competing Interests: Te authors declare that they have no competing interests. Publisher's note: Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional afliations. Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Cre- ative Commons license, and indicate if changes were made. Te images or other third party material in this article are included in the article’s Creative Commons license, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons license and your intended use is not per- mitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this license, visit http://creativecommons.org/licenses/by/4.0/.

© Te Author(s) 2017

SCientifiC RepOrTs | 7:13641 | DOI:10.1038/s41598-017-13294-3 16