Skip to main content

Theory and Modern Applications

Global dynamics of a delayed chemostat model with harvest by impulsive flocculant input

Abstract

A mathematical model describing continuous microbial culture and harvest in a chemostat, incorporating a control strategy and defined by impulsive differential equations, is presented and investigated. Theoretical results indicate that the model has a microbe-extinction periodic solution, which is globally attractive if the threshold \(R_{1}\) is less than unity, and the model is permanent if the threshold \(R_{2}\) is greater than unity. Further, we consider the control strategy under time delay and periodical impulsive effect. Analysis shows that continuous microbial culture and harvest process can be implemented by adjusting time delay, impulsive period or input amount of flocculant. Finally, we give an example with numerical simulations to illustrate the control strategy.

1 Introduction and model formulation

A chemostat is a classical bioreactor for microbes culture and has been widely applied in the field of microbiology and bioengineering [1, 2]. The chemostat model has attracted the attention of many scholars since it was introduced by Monod in 1942 [3]. These models include mathematical models [4–16] and experimental models [17–19]. A simple chemostat can be designed by using a pump or an overflow system (see Figure 1), by which the volume of the chemostat can be controlled either [20].

Figure 1
figure 1

Two types of chemostat. (a) Type I chemostat. (b) Type II chemostat.

It was found that some microbes can produce flocculation under the action of flocculating agents. This phenomenon makes it possible to harvest microbes by flocculating agents [21–32]. Recently, based on a classical simple chemostat model in which a microbial species consumes a single growth limiting substrate [3, 8], Tai et al. [33] have proposed a delayed differential equations (DDEs) model to describe the process of microbial continuous culture and harvest as follows:

$$ \textstyle\begin{cases} {\frac{dS(t)}{dt}=D(S_{0}-S(t))-h_{1}x(t)S(t)},\\ {\frac{dx(t)}{dt}=hx(t-\tau)S(t-\tau )-Dx(t)-h_{2}x(t)F(t)},\\ {\frac{dF(t)}{dt}=D(F_{0}-F(t))-h_{3}x(t)F(t)}, \end{cases} $$
(1)

where \(S(t)\), \(x(t)\) and \(F(t)\) represent the concentration of the substrate, microbia and flocculating agent at time t, respectively. D represents the velocity of medium, \(S_{0}\) and \(F_{0}\) represent the input concentration of substrate and flocculating agent, respectively. \(h_{1}\) and h represent the consumption of medium and the yield of microbia, respectively. \(h_{3}\) is the loss rate of flocculant. τ represents the time involved converting nutrient into the microbia [34–45]. The authors found that the model can produce backward bifurcation and complex dynamics. By establishing analytic thresholds for the existence of backward bifurcation, they analyzed the local stability of the equilibria.

In model (1), the flocculating agent is assumed to be added into the chemostat continuously. While in practice, by considering resource savings and the growth cycle of microorganisms, the flocculating agent can be periodically added into the chemostat at some fixed moment. Thus we perfect the chemostat system by adding input channel of flocculants and output channel of flocculation by two pumps, respectively (see Figure 2). It can be regulated through inputting flocculant to flocculate microbia according to the concentration of microbia. This process can be described by impulsive differential equations (IDEs). Impulsive differential equations, on the one hand, can fully reflect the actual control situation; on the other hand, they can guide the operator to implement the impulsive control strategy conveniently and accurately [46–51]. Thus, we propose a new continuous culture chemostat model with time delay and impulsive harvest as follows:

$$ \textstyle\begin{cases} \left . \textstyle\begin{array}{l} {\frac{dS(t)}{dt}=D(S_{0}-S(t))-h_{1}x(t)S(t)},\\ {\frac{dx(t)}{dt}=e^{-D\tau}hx(t-\tau)S(t-\tau )-Dx(t)-h_{2}x(t)F(t)},\\ {\frac{dF(t)}{dt}=-DF(t)-h_{3}x(t)F(t)}, \end{array}\displaystyle \right \} t\neq nT,\\ \left . \textstyle\begin{array}{l} S(t^{+})=S(t), \\ x(t^{+})=x(t), \\ F(t^{+})=F(t)+\gamma F_{0}, \end{array}\displaystyle \right \}t=nT, \end{cases} $$
(2)

where T is the period of the impulsive effect, \(\gamma F_{0}\) is the input amount of flocculant at every impulsive period T. \(F(nT^{+})= \lim _{t\to nT^{+}}F(t)\), and \(F(t)\) is left continuous at \(t=nT\), i.e., \(F(nT)= \lim _{t\to nT^{-}}F(t)\), \(S(t)\), \(x(t)\) are continuous for all \(t\geq0\), the details can be seen in [52, 53].

Figure 2
figure 2

The modified chemostat.

Let \(C_{+}=C([-\tau,0],R^{3}_{+})\) be the Banach space, \(\psi=(\psi _{1}(s),\psi_{2}(s),\psi_{3}(s))^{T}\), \(\psi_{i}(\theta)\geq0\) (\(-\tau\leq \theta\leq0\), \(i=1,2,3\)) the initial conditions are given as

$$ \begin{aligned}&S(\theta)=\psi_{1}(\theta), \qquad x(\theta)= \psi_{2}(\theta), \qquad F(\theta)=\psi _{3}(\theta), \\ &\psi\in C_{+},\qquad \psi_{i}(0)>0 \quad(i=1,2,3). \end{aligned}$$
(3)

The rest of the paper is organized as follows. In Section 2, we briefly introduce some concepts and fundamental results, which are necessary for future discussion. In Section 3, we focus our attention on the global property of system (2), including the existence, global attractivity of the microbe-extinction periodic solution and the permanence of system (2). In Section 4, we give the threshold of key parameters of system (2) and discuss the control strategy. We finally give a conclusion and numerical simulations in Section 5, from which it can be seen that all simulations agree with the theoretical results.

2 Preliminaries

In this section, we give some useful lemmas.

Let \(f=(f_{1}, f_{2}, f_{3})^{T}\) be the map defined by the right-hand side of the anterior three equations of system (2). Let \(R_{+} = [0,\infty)\), \(R^{3}_{+}=\{x\in R^{3}: x \geq0\}\), \(\Omega=\operatorname{int}R^{3}_{+}\). Let \(U:R_{+}\times R_{+}^{3}\rightarrow R_{+}\). If U satisfies the following conditions: (1) V is continuous in \(((n-1)T,nT]\times R_{+}^{3}\), \(n\in N\), and for each \(x\in R_{+}^{3}\), \(\lim _{(t,z)\to((n-1)T^{+},x)}U(t,z)=U((n-1)T,x)\) and \(\lim _{(t,z)\to(nT^{+},x)}U(t,z)=U(nT^{+},x)\) exists; (2) U is locally Lipschitzian in x. Then U is said to belong to class \(U_{0}\).

Lemma 2.1

[52, 53]

Let \(U: R_{+}\times R^{3}_{+}\rightarrow R_{+}\), \(H: R_{+}\times R_{+}\rightarrow R\) and \(U\in U_{0}\). Assume that

$$ \textstyle\begin{cases} D^{+}U(t, w(t))\leq(\geq)\ H(t, U(t, w(t))),& t\neq n\omega,\\ U(t, w(t)^{+})\leq(\geq)\ \Upsilon_{n}(U(t, w(t))),& t=n\omega, \end{cases} $$
(4)

here H is continuous in \((n\omega, (n+1)\omega] \times R_{+}\) and \(\forall x\in R_{+}\), \(n \in N\), \(\lim _{(t,y)\to((n\omega )^{+},x)} H(t, y) = H((n\omega)^{+}, x)\) exist; \(\Upsilon_{n}: R_{+}\to R_{+}\) is nondecreasing. Let \(r(t)=r(t, 0, u_{0})\) be the maximal (minimal) solution of the scalar impulsive differential equation

$$ \textstyle\begin{cases} u'=H(t, u), & t\neq n\omega,\\ u(t^{+})=\Upsilon_{n}(u(t)), &t=n\omega, \end{cases} $$
(5)

existing on \([0, \infty)\). Then \(U(0^{+}, w_{0})\leq(\geq)\ u_{0}\) implies that \(U(t, w(t)) \leq(\geq)\ r(t)\), \(t \geq0\), where \(\omega(t)=\omega(t, 0, w_{0})\) is any solution of (4) existing on \([0, \infty)\).

Lemma 2.2

[54]

Let \(q_{1}\), \(q_{2}\), Ï„ be all positive constants and \(z(t)>0\) for \(t\in[-\tau,0]\). Consider the following delay differential equation:

$$\frac{dz(t)}{dt}=q_{1}z(t-\tau)-q_{2}z(t), $$

then

  1. (i)

    if \(q_{1}< q_{2}\), then \(\lim _{t\to\infty}z(t)=0\);

  2. (ii)

    if \(q_{1}>q_{2}\), then \(\lim _{t\to\infty }z(t)=\infty\).

Lemma 2.3

[52, 53]

Consider the following impulse differential inequalities:

$$ \textstyle\begin{cases} u'(t)\leq(\geq)\ a(t)u(t)+c(t), & t\neq t_{k},\\ u(t^{+}_{k})\leq(\geq)\ b_{k}u(t_{k})+d_{k}, & t=t_{k}, k\in N, \end{cases} $$

where \(a(t),c(t)\in C(R_{+},R)\), \(b_{k}\geq0\), and \(d_{k}\) are constants. Assume

\((A_{0})\) :

the sequence \(\{t_{k}\}\) satisfies \(0\leq t_{0}< t_{1}< t_{2}<\cdots\), with \(\lim_{t\rightarrow \infty}t_{k}=\infty\);

\((A_{1})\) :

\(u\in PC'(R_{+},R)\) and \(u(t)\) is left-continuous at \(t_{k}\), \(k\in N\). Then

$$\begin{aligned} u(t) \leq (\geq)&u(t_{0}) \prod_{t_{0}< t_{k}< t}d_{k} \exp \biggl( \int_{t_{0}}^{t}a(s)\,ds \biggr)+ \sum _{t_{0}< t_{k}< t} \biggl(\prod_{t_{k}< t_{j}< t}d_{j} \exp \biggl( \int_{t_{k}}^{t}a(s)\,ds \biggr) \biggr)d_{k} \\ &{}+ \int_{t_{0}}^{t}\prod_{s< t_{k}< t}b_{k} \exp \biggl( \int_{s}^{t}a(\theta )\,d\theta \biggr)c(s)\,ds, \quad t \geq t_{0}. \end{aligned}$$

Lemma 2.4

[43]

Consider the following impulsive differential system:

$$ \textstyle\begin{cases} {\frac{d\hbar(t)}{dt}=r_{1}-r_{2}\hbar(t)},& t\neq nT,\\ {\hbar}(t^{+})={\hbar}(t)+\mu, & t=nT, \end{cases} $$
(6)

for each solution \(\hbar(t)\) of (6), \(\hbar(t) \rightarrow\hbar^{*}(t)\) as \(t\rightarrow\infty\), where \(\hbar^{*}(t)=\frac{r_{1}}{r_{2}}+\frac{\mu e^{-r_{2}(t-nT)}}{1-e^{-r_{2}T}}\) for \(t\in(nT,(n+1)T]\).

Lemma 2.5

There exist constants \(M_{1},M_{2},M_{3}>0\) such that \(S(t)\leq M_{1}\), \(x(t)\leq M_{2}\), \(F(t)\leq M_{3}\) for each solution of (2) with all t large enough.

Proof

Firstly, from the third and sixth equation of system (2), we have

$$ \textstyle\begin{cases} {\frac{dF(t)}{dt}\leq-FS(t)}, & t\neq nT,\\ F(t^{+})=F(t)+\gamma F_{0},& t=nT. \end{cases} $$

By Lemma 2.3, we have \(F(t)\le\gamma F_{0}\frac {e^{DT}}{e^{DT}-1}+\varepsilon_{2}\) for t large enough.

Let \(V(t)=e^{-D\tau}\frac{h}{h_{1}}S(t-\tau)+x(t)\). It is clear that \(V\in U_{0}\). Calculating the upper right derivative of \(V(t)\) along a solution of system (2), one can get

$$ \frac{dV(t)}{dt} \leq e^{-D\tau}\frac{hDS_{0}}{h_{1}}-DV(t), $$

then by Lemma 2.3, we have \(\limsup _{t\to\infty} V(t)\le e^{-D\tau}\frac{hS_{0}}{h_{1}}\), so \(V(t)\) is ultimately bounded. Thus, \(S(t)\) and \(x(t)\) are ultimately bounded and \(\limsup _{t\to\infty}S(t)\le S_{0}\), \(\limsup _{t\to\infty}x(t)\le e^{-D\tau}\frac{hS_{0}}{h_{1}}\). Let \(M_{1}=S_{0}+\varepsilon\), \(M_{2}=e^{-D\tau}\frac{hS_{0}}{h_{1}}+\varepsilon\), \(M_{3}=\gamma F_{0}\frac{e^{DT}}{e^{DT}-1}+\varepsilon\), we have \(S(t)\leq M_{1}\), \(x(t)\leq M_{2}\), \(F(t)\le M_{3}\) for t large enough. The proof is completed. □

3 Global dynamical analysis for system (2)

In this section, we discuss the global dynamics of model (2), including the existence and global attractivity of the microbe-extinction periodic solution and the permanence.

3.1 Existence and global attractivity of the microbe-extinction periodic solution

Microbe-extinction solution describes that microbes are eventually absent from system (2), thus we let \(x(t)=0\) in system (2), then system (2) changes to the following system:

$$ \textstyle\begin{cases} \left . \textstyle\begin{array}{l} {\frac{dS(t)}{dt}=D(S_{0}-S(t))},\\ {\frac{dF(t)}{dt}=-DF(t)}, \end{array}\displaystyle \right \} t\neq nT,\\ \left . \textstyle\begin{array}{l} S(t^{+})=S(t), \\ F(t^{+})=F(t)+\gamma F_{0}, \end{array}\displaystyle \right \}t=nT. \end{cases} $$
(7)

Note that the variates \(S(t)\) and \(F(t)\) are independent of each other in system (7). Thus, by Lemma 2.4, we obtain that system (7) has a unique positive T-periodic solution \((S^{*}(t), F^{*}(t))\) and for each solution \((S(t), F(t))\) of system (7), \(S(t) \rightarrow S^{*}(t)\) and \(F(t) \rightarrow F^{*}(t)\) as \(t\rightarrow\infty\), where

$$ \textstyle\begin{cases} S^{*}(t)=S_{0},\\ F^{*}(t)=\frac{\gamma F_{0} e^{-D(t-nT)}}{1-e^{-DT}}. \end{cases} $$
(8)

Therefore, we have the existence theorem for system (2).

Theorem 3.1

System (2) has a microbe-extinction periodic solution \((S_{0}, 0, F^{*}(t))\).

Denote

$$ R_{1}=\frac{he^{-D\tau}S_{0}}{D+h_{2}\frac{\gamma F_{0}e^{-(D+h_{3}M_{2})T}}{1-e^{-(D+h_{3}M_{2})T}}}. $$

We have the following theorem about the attractivity of the microbe-extinction periodic solution of system (2).

Theorem 3.2

If \(R_{1}<1\), then the microbe-extinction periodic solution \((S_{0}, 0, F^{*}(t))\) of system (2) is globally attractive.

Proof

Let \((S(t), x(t), F(t))\) be any solution of system (2) satisfying initial condition (3). Since \(R_{1}<1\), one can choose \(\varepsilon_{1}, \varepsilon_{2}>0\) such that

$$ he^{-D\tau}(S_{0}+\varepsilon_{1})< D+h_{2} \biggl(\frac{\gamma F_{0}e^{-(D+h_{3}M_{2})T}}{1-e^{-(D+h_{3}M_{2})T}}-\varepsilon_{2} \biggr). $$
(9)

By the first equation of system (2), we have

$$ \frac{dS(t)}{dt}\leq D\bigl(S_{0}-S(t)\bigr). $$

According to Lemma 2.3, we have

$$ \limsup_{t\rightarrow\infty} S(t)\leq S_{0}. $$

Hence, there exists \(n_{1}\in N^{+}\) such that

$$ S(t)\leq S_{0}+\varepsilon_{1} $$
(10)

for all \(t\geq n_{1}T\), where \(\varepsilon_{1}\) is an arbitrarily small positive constant.

By the third and sixth equations of system (2), we have

$$ \textstyle\begin{cases} \frac{dF(t)}{dt}\geq-(D+h_{3}M_{2})F(t),& t\neq nT,\\ F(t^{+})=F(t)+\gamma F_{0},& t=nT, \end{cases} $$
(11)

then consider the following impulsive differential system:

$$ \textstyle\begin{cases} \frac{dq(t)}{dt}=-(D+h_{3}M_{2})q(t),& t\neq nT,\\ q(t^{+})=q(t)+\gamma F_{0},& t=nT,\\ q(0^{+})=F(0^{+}). \end{cases} $$
(12)

Then, by using Lemma 2.1, we have \(F(t)\geq q(t)\) and \(q(t)\rightarrow q^{*}(t)\) as \(t\rightarrow\infty\), \(q^{*}(t)\) is the periodic solution of (12), where \(q^{*}(t)=\frac {\gamma F_{0} e^{-(D+h_{3}M_{2})(t-nT)}}{1-e^{-(D+h_{3}M_{2})T}}\), \(nT< t\leq (n+1)T\). By Lemma 2.4, we have that \(q^{*}(t)\) is globally asymptotically stable. Hence there exists \(n_{2}\in N^{+}\) such that

$$ F(t)\geq{q(t)> q^{*}(t)}-\varepsilon_{2}> \frac{\gamma F_{0} e^{-(D+h_{3}M_{2})T}}{1-e^{-(D+h_{3}M_{2})T}}-\varepsilon_{2} $$
(13)

for all \(t\geq n_{2}T\), where \(\varepsilon_{2}\) is an arbitrarily small positive constant.

From the second equation, (10) and (13), there exists a positive integer \(n_{3}>\max\{n_{1},n_{2}\}\), for \(t>n_{3}T+\tau\), we have

$$ \frac{dx(t)}{dt}\leq he^{-D\tau}(S_{0}+\varepsilon_{1})x(t- \tau)- \biggl(D+h_{2}\frac{\gamma F_{0}e^{-(D+h_{3}M_{2})T}}{1-e^{-(D+h_{3}M_{2})T}}-\varepsilon _{2} \biggr)x(t). $$

Consider the following delay differential equation:

$$ \frac{dy(t)}{dt}= he^{-D\tau}(S_{0}+\varepsilon_{1})y(t- \tau)- \biggl(D+h_{2} \biggl(\frac{\gamma F_{0}e^{-(D+h_{3}M_{2})T}}{1-e^{-(D+h_{3}M_{2})T}}-\varepsilon_{2} \biggr) \biggr)y(t). $$

Since (9) holds, by Lemma 2.2, we get \({\lim } _{t\to\infty}y(t)= 0\). Notice that for all \(\theta\in[-\tau, 0]\), \(x(\theta) = y(\theta) = \psi_{2}(\theta)>0\) holds. By the comparison theorem in differential equation and the positivity of solution (with \(x(t)\geq0\)), we obtain \(\lim _{t\to\infty}x(t)= 0\).

Next, we will prove \(\lim _{t\to\infty}S(t)=S_{0}\) and \(\lim _{t\to\infty}F(t)=F^{*}(t)\). We assume that \(0< x(t)<\varepsilon\) holds for all \(t\geq0\) in the following discussion without loss of generality. One the one hand, by the first equation of system (2), one gets

$$ \frac{dS(t)}{dt}\geq DS_{0}-(D+\varepsilon h_{1})S(t). $$

Then, we have \(\liminf _{t\to\infty} S(t)\geq S_{0}\frac {D}{D+\varepsilon h_{1}}\). Thus, there exists \(T>0\) such that for any \(\varepsilon_{1}>0\),

$$ S(t)\geq S_{0}\frac{D}{D+\varepsilon h_{1}}-\varepsilon_{1} $$
(14)

for \(t>T\). Let \(\varepsilon\rightarrow0\), from (10) and (14) we have

$$ S_{0}-\varepsilon_{1} < S(t)< S_{0} + \varepsilon_{1} $$

for t large enough, then we have \(\lim _{t\to\infty}S(t)=S_{0}\).

On the other hand, from the third and sixth equations of system (2), we have

$$ \textstyle\begin{cases} \frac{dF(t)}{dt}\geq-(D+\varepsilon h_{3})F(t),& t\neq nT,\\ F(t^{+})=F(t)+\gamma F_{0},& t=nT. \end{cases} $$

Then we get the following comparison system:

$$ \textstyle\begin{cases} \frac{dw_{1}(t)}{dt}=-(D+\varepsilon h_{3})w_{1}(t),& t\neq nT,\\ w_{1}(t^{+})=w_{1}(t)+\gamma F_{0},& t=nT,\\ w_{1}(0^{+})=F(0^{+}). \end{cases} $$
(15)

By Lemma 2.4, system (15) has a globally asymptotically stable positive periodic solution \(w_{1}^{*}(t)\), where \(w_{1}^{*}(t)=\frac{\gamma F_{0} e^{-(D+\varepsilon h_{3})(t-nT)}}{1-e^{-(D+\varepsilon h_{3})T}}\). Thus, by Lemma 2.1, we have \(F(t)\geq w_{1}(t)\) and \(w_{1}(t)\rightarrow w_{1}^{*}(t)\) as \(t\rightarrow\infty\). Therefore, there exists \(T'>0\) such that for any \(\varepsilon_{2}>0\),

$$ F(t)\geq w_{1}^{*}(t)-\varepsilon_{2} $$
(16)

for \(t>T'\). From the third and sixth equations of (2), one gets

$$ \textstyle\begin{cases} \frac{dF(t)}{dt}\leq-DF(t),& t\neq nT,\\ F(t^{+})=F(t)+\gamma F_{0},& t=nT. \end{cases} $$

Consider the following comparison system:

$$ \textstyle\begin{cases} \frac{dw_{2}(t)}{dt}=-Dw_{2}(t),& t\neq nT,\\ w_{2}(t^{+})=w_{2}(t)+\gamma F_{0},& t=nT,\\ w_{2}(0^{+})=F(0^{+}), \end{cases} $$
(17)

then we have \(F(t)\leq w_{2}(t)\) and \(w_{2}(t)\rightarrow F^{*}(t)\). Then, for any \(\varepsilon_{2} >0\), there exists \(T''>0\) such that

$$ F(t)\leq F^{*}(t)-\varepsilon_{2} $$
(18)

for \(t>T''\). Thus, by (16) and (18), for \(t>\max\{T',T''\}\), we have

$$ w_{1}^{*}(t)-\varepsilon_{2}\leq F(t)\leq F^{*}(t)- \varepsilon_{2}. $$

Let \(\varepsilon\rightarrow0\), we have

$$ F^{*}(t)-\varepsilon_{2} < F(t)< F^{*}(t)+ \varepsilon_{2} $$

for t large enough, thus we get \(\lim _{t\to\infty }F(t)=F_{0}\). This completes the proof. □

3.2 Permanence

In this section, we prove that system (2) is persistent for \(R_{2}>1\). Firstly, we give the following lemma supporting our main conclusion. Denote

$$ R_{2}=\frac{he^{-D\tau}S_{0}}{D+h_{2}\frac{\gamma F_{0}}{1-e^{-DT}}}. $$

Lemma 3.1

If \(R_{2}>1\), then there exists a constant \(m_{4}>0\) such that

$$ \liminf _{t\to\infty}x(t)\geq\min \biggl\{ \frac {m_{4}}{2},m_{4}e^{-(D+h_{2}\varrho_{2})\tau} \biggr\} =m_{2}. $$

Proof

Let \(X(t)=(S(t), x(t), F(t))\) be any positive solution of system (2) with initial condition (3). We rewrite the second equation of system (2) as follows:

$$ \frac{dx(t)}{dt}=\bigl(e^{-D\tau }hS(t)-D-h_{2}F(t) \bigr)x(t)-e^{-D\tau}h \int_{t-\tau}^{t} x(\sigma)S(\sigma )\,d\sigma. $$

Define

$$ G(t)=x(t)+e^{-D\tau}h \int_{t-\tau}^{t} x(\sigma)S(\sigma )\,d\sigma. $$

Calculating the derivative of \(G(t)\) along the solution of (2) yields

$$ \frac{dG(t)}{dt}= \bigl(e^{-D\tau}hS(t)-D-h_{2}F(t) \bigr)x(t). $$
(19)

Let \({m_{4}=\frac{(R_{2}-1)D}{(R_{2}+1)h_{1}}}\). Since \(R_{2}>1\), it is clear that \(m_{4}>0\). For \(m_{4}\), one can choose \(\varepsilon _{1},\varepsilon_{2}>0\) small enough such that

$$ \frac{e^{-D\tau}h\varrho_{1}}{D+h_{2}\varrho_{2}}>1, $$
(20)

where \(\varrho_{1}=\frac{DS_{0}}{D+h_{1}m_{4}}-\varepsilon_{1}\), \(\varrho_{2}=\frac {\gamma F_{0}}{1-e^{-DT}}+\varepsilon_{2}\). Then, for any positive constant \(t_{0}\) and for all \(t\geq t_{0}\), we claim that the inequality \(x(t)< m_{4}\) cannot hold. Otherwise, there must exist a positive constant \(t_{0}\) such that \(x(t)< m_{4}\) for all \(t\geq t_{0}\). The first equation of system (2) leads to

$$ \frac{dS(t)}{dt}\geq DS_{0}-(D+h_{1}m_{4})S(t). $$

By Lemma 2.3, there exists such \(T_{1} >t_{0}+\tau\) for \(t>T_{1}\) that

$$ S(t)>\frac{DS_{0}}{D+h_{1}m_{4}}-\varepsilon_{1}= \varrho_{1}. $$
(21)

From (18), there exists such \(T_{2} >t_{0}+\tau\) for \(t\geq T_{2}\) that

$$ F(t)< F^{*}(t)+ \varepsilon_{2}=\frac{\gamma F_{0} e^{-D(t-nT)}}{1-e^{-DT}}+ \varepsilon_{2}< \frac{\gamma F_{0}}{1-e^{-DT}}+\varepsilon_{2}= \varrho_{2}. $$
(22)

Thus by (21), (22) and (19), for \(t>T_{3}=\max \{T_{1},T_{2}\}\), one gets

$$ \frac{dG(t)}{dt}\geq(D+h_{2}\varrho) \biggl( \frac{e^{-D\tau}h\varrho _{1}}{D+h_{2}\varrho_{2}}-1 \biggr)x(t). $$
(23)

Let

$$ x_{l}=\min_{t\in[T_{1},T_{1}+\tau]}x(t). $$

We can prove that \(x(t)\geq x_{l}\) for all \(t\geq T_{3}\). Otherwise, there exists a constant \(T_{4}\geq0\) such that \(x(t)\geq x_{l}\) for \(t\in[T_{3},T_{3}+\tau+T_{4}]\), \(x(T_{3}+\tau +T_{4})=x_{l}\) and \(\dot{x}(T_{3}+\tau+T_{4})\leq0\). Then from the second equation of (2), (20) and (23), we have

$$\begin{aligned} \dot{x}(T_{3}+\tau+T_{4}) >& \bigl(e^{-D\tau}h \varrho_{1}-D-h_{2}\varrho_{2} \bigr)x_{l} \\ =&(D+h_{2}\varrho_{2}) \biggl(\frac{e^{-D\tau}h\varrho_{1}}{D+h_{2}\varrho _{2}}-1 \biggr)x_{l} \\ >&0, \end{aligned}$$

which is a contradiction. Hence \(x(t)\geq x_{l}>0\) for all \(t\geq T_{3}\). Inequality (23) implies

$$ \frac{dG(t)}{dt}\geq(D+h_{2}\varrho) \biggl( \frac{e^{-D\tau}h\varrho _{1}}{D+h_{2}\varrho_{2}}-1 \biggr)x_{l}>0, $$
(24)

which implies \(G(t)\rightarrow+\infty\) as \(t\rightarrow+\infty\). This is a contradiction to \(G(t)\leq M(1+M\tau e^{-DM})\) for t large enough. Therefore, for any positive constant \(t_{0}\), the inequality \(x(t)< m_{4}\) cannot hold for all \(t\geq t_{0}\).

Step II: From Step I, we only need to consider:

  1. (i)

    \(x(t)>m_{4}\) for all t large enough;

  2. (ii)

    \(x(t)\) oscillates about \(m_{4}\) for all large t.

However, case (i) is obviously the result of this theorem, so we only need to consider case (ii), in which we shall show that \(x(t)>m_{2}\) for all large t, where

$$ m_{2}=\min \biggl\{ \frac{m_{4}}{2},m_{4}e^{-(D+h_{2}\varrho _{2})\tau} \biggr\} . $$

First, we notice that there would be two positive arbitrarily big constants t̄, φ such that \(x(t)< m_{4}\) for \(\bar{t}< t<\bar {t}+\varphi\) and \(x(\bar{t})=x(\bar{t}+\varphi)=m_{4}\). Second, there exists a constant \(0< T_{5}<\tau\) such that \(x(t)>\frac {m_{4}}{2}\) for all \(\bar{t}\leq t\leq\bar{t}+T_{5}\). Because \(x(t)\) is not affected by impulses and, moreover, \(x(t)\) is bound and continuous, then we conclude that \(T_{5}\) is independent of the choice of t̄. Next, according to the position of φ, \(T_{5}\), τ, there will be three cases we should discuss.

Case ii(a): \(\varphi\leq T_{5}\), obviously our aim is obtained.

Case ii(b): \(T_{5}<\varphi\leq\tau\). By (22), the second equation of (2) implies

$$ \dot{x}(t)\geq-(D+h_{2}\varrho_{2})x(t) $$

for \(\bar{t}< t \leq\bar{t}+\varphi<\bar{t}+\tau\). Then we have

$$ x(t)\geq x(\bar{t})e^{-(D+h_{2}\varrho_{2})\tau} $$

for \(\bar{t}< t \leq\bar{t}+\varphi \leq\bar{t}+\tau\), notice \(x(\bar{t})=m_{4}\), one can get

$$ x(t)\geq m_{4}e^{-(D+h_{2}\varrho_{2})\tau} $$

for \(\bar{t}< t \leq\bar{t}+\varphi \leq\bar{t}+\tau\). Thus we have \(x(t)\geq m_{2}\) for \(\bar{t}< t \leq\bar{t}+\varphi\).

Case ii(c): \(\varphi\geq\tau\). We have proved \(x(t)\geq m_{2}\) for \(\bar{t}< t \leq\bar{t}+\tau\). For \(\bar{t}+\tau\leq t\leq\bar{t}+\varphi \), we can analyze and prove \(x(t)\geq m_{2}\) as the proof for the above claim. Because of the arbitrariness of interval \([\bar{t},\bar {t}+\varphi]\) and because t̄ is an arbitrarily big constant, we have that \(x(t)\geq m_{2}\) holds for t large enough. Finally, notice that the choice of \(m_{2}\) is independent of the positive solution of (2), which satisfies that \(x(t)\geq m_{2}\) for t large enough. This completes the proof of Lemma 2.4. □

Theorem 3.3

For \(R_{2}>1\), then system (2) will be permanent.

Proof

Let \(X(t)=(S(t), x(t), F(t))\) be any positive solution of system (2) with initial condition (3). From the first equation of system (2), we get

$$ \frac{dS(t)}{dt}\geq DS_{0}-(D+h_{1}M)S(t), $$

then we have

$$ \liminf_{t\to\infty}S(t)\geq\frac{DS_{0}}{D+h_{1}M}. $$

Thus there exists a constant ε small enough such that

$$ S(t)>\frac{DS_{0}}{D+h_{1}M}-\varepsilon=m_{1}>0 $$

for t large enough. And from (13) we have

$$ F(t)\geq z(t)> z^{*}(t)-\varepsilon>\frac{\gamma F_{0} e^{-(D+h_{3}M)T}}{1-e^{-(D+h_{3}M)T}}=m_{3}>0 $$
(25)

for t large enough.

Set

$$ D=\bigl\{ (S,x,F)\in R^{3}_{+}|m_{1}\leq S\leq M_{1},m_{2}\leq x\leq M_{2},m_{3}\leq F\leq M_{3}\bigr\} . $$

Thus D is a bounded compact region and every solution of system (2) will eventually enter and remain in region D, then system (2) is permanent. The proof of Theorem 3.3 is completed. □

4 Control strategy of continuous microbial culture and harvest

In Section 3, we obtain the threshold values \(R_{1}\) and \(R_{2}\) associated with microbial extinction and existence. Next, we discuss the control strategy of continuous microbial culture and harvest by analyzing the key parameters of the threshold.

Denote

$$\begin{aligned}& T^{*}=\frac{1}{D+h_{3}M_{2}}\ln\frac{he^{-D\tau }S_{0}+h_{2}\gamma F_{0}-D}{he^{-D\tau}S_{0}-D}, \\& F_{0}^{*}=\frac{he^{-D\tau }S_{0}-D}{h_{2}\frac{\gamma e^{-(D+h_{3}M_{2})T}}{1-e^{-(D+h_{3}M_{2})T}}}, \\& \tau ^{*}=\frac{1}{D}\ln\frac{hS_{0}}{D+h_{2}\frac{\gamma F_{0}e^{-(D+h_{3}M_{2})T}}{1-e^{-(D+h_{3}M_{2})T}}}, \\& T_{*}=\frac{1}{D}\ln\frac{he^{-D\tau}S_{0}+h_{2}\gamma F_{0}-D}{he^{-D\tau }S_{0}-D}, \\& F_{0*}=\frac{(he^{-D\tau}S_{0}-D)(1-e^{-DT})}{h_{2}\gamma}, \\& \tau _{*}=\frac{1}{D}\ln\frac{hS_{0}}{D+h_{2}\frac{\gamma F_{0}}{1-e^{-DT}}}. \end{aligned}$$

According to Theorem 3.2, we have that if \(T< T^{*}\) or \(F_{0}>F_{0}^{*}\) or \(\tau>\tau^{*}\), the microbe-extinction periodic solution \((S_{0}, 0, F^{*}(t))\) is globally attractive. That means the microbial continuous cultivation and harvest have failed. And from Theorem 3.3, we know that if \(T>T_{*}\) or \(F_{0}< F_{0*}\) or \(\tau<\tau_{*}\), then system (2) is permanent. That means we can achieve the process of microbial cultivation by increasing the time interval, or reducing the input amount of flocculant, or shortening the growth delay.

5 Discussion and numerical simulations

In this paper, to achieve the continuous microbial culture and harvest, we improve the classic chemostat model and propose a new chemostat model with time delay and periodical flocculant input. Our main aim is to investigate the control strategy of continuous microbial culture and harvest. By using the theory of impulsive delayed differential equations, global properties of the system are discussed. We prove that if \(R_{1}<1\), then the microbe will be eventually extinct, and if \(R_{2}>1\), the microbe species is permanent. Based on the threshold values associated with microbial extinction and existence, we consider the control strategy. Results show that we can culture microbia continuously and harvest microbia many times by adjusting the time interval (T) or the input amount of flocculant (\(\gamma F_{0}\)), or the time delay (Ï„).

Next we will verify the effectiveness of control strategy by an example and some numerical simulations. Let \(D=0.3\), \(S_{0}=2.8\), \(h_{1}=0.4\), \(h=0.3\), \(h_{2}=0.15\), \(h_{3}=0.02\), \(\gamma=1\), and let the initial value be \((1,1,0)\). We get the following system:

$$ \textstyle\begin{cases} \left . \textstyle\begin{array}{l} \frac{dS(t)}{dt}=0.3(2.8-S(t))-0.4x(t)S(t),\\ \frac{dx(t)}{dt}=0.3e^{-0.3\tau}x(t-\tau)S(t-\tau )-0.3x(t)-0.15x(t)F(t),\\ \frac{dF(t)}{dt}=-0.3F(t)-0.02x(t)F(t), \end{array}\displaystyle \right \} t\neq nT,\\ \left . \textstyle\begin{array}{l} S(t^{+})=S(t), \\ x(t^{+})=x(t), \\ F(t^{+})=F(t)+\gamma F_{0}, \end{array}\displaystyle \right \}t=nT. \end{cases} $$
(26)

To investigate the effects of key parameters on the system, we assume \(\tau=2\), \(T=2\), \(F_{0}=1\). By simple calculation, we have \(R_{1}=0.9910<1\). Figure 3 shows that the microbe-eradication periodic solution \((S_{0}, 0, F^{*}(t))\) is globally attractive. That means microbial continuous cultivation and harvest have failed because the microbe will be eventually extinct. To achieve microbial continuous cultivation and harvest, we can take three kinds of control strategies.

  1. (i)

    We can reduce the input of flocculant \(F_{0}\) (from 1 to 0.4). By calculating, we have \(R_{2}=1.0647>1\). According to Theorem 3.3, system (26) is permanent. A lower amount of flocculant can increase population microbia in the medium so that microbe can be cultured continuously in the chemostat system. Figure 4 shows that system (26) is permanent.

  2. (ii)

    We can increase the time travail T (from 2 to 10). Longer time travail can decrease input mount of flocculant indirectly and increase population microbia in the medium, which makes microbia cultured and harvested continuously. By calculating, we have \(R_{2}=1.0069>1\). According to Theorem 3.3, system (26) is permanent (see Figure 5). Figure 5 also shows that system (26) has an asymptotically stable periodic solution.

  3. (iii)

    We can decrease time delay τ (from 2 to 0.5) by some biotechnology and biological engineering. Reduction of growth time delay makes the microbial growth cycle shorter, which is beneficial for microbial continuous cultivation. Figure 6 shows that system (26) is permanent. Moreover, system (26) has an asymptotically stable periodic solution (where \(R_{2}=1.1432>1\)). Detailed parameter values, thresholds and states of system (26), please see Table 1.

    Table 1 Values of parameters, threshold and state of the system
Figure 3
figure 3

Dynamical behavior of solutions of system ( 26 ) with \(\pmb{R_{1}=0.9910<1}\) . (a) Time series of \(S(t)\), \(x(t)\), \(F(t)\). (b) Three-dimensional phase diagrams of \(S(t)\), \(x(t)\), \(F(t)\). (c) Phase diagrams of \(x(t)\), \(F(t)\). (d) Phase diagrams of \(x(t)\), \(S(t)\).

Figure 4
figure 4

Dynamical behavior of solutions of system ( 26 ) with \(\pmb{R_{2}=1.0647>1}\) . (a) Time series of \(S(t)\), \(x(t)\), \(F(t)\). (b) Three-dimensional phase diagrams of \(S(t)\), \(x(t)\), \(F(t)\). (c) Phase diagrams of \(x(t)\), \(F(t)\). (d) Phase diagrams of \(x(t)\), \(S(t)\).

Figure 5
figure 5

Dynamical behavior of solutions of system ( 26 ) with \(\pmb{R_{2}=1.0069>1}\) . (a) Time series of \(S(t)\), \(x(t)\), \(F(t)\). (b) Three-dimensional phase diagrams of \(S(t)\), \(x(t)\), \(F(t)\). (c) Phase diagrams of \(x(t)\), \(F(t)\). (d) Phase diagrams of \(x(t)\), \(S(t)\).

Figure 6
figure 6

Dynamical behavior of solutions of system ( 26 ) with \(\pmb{R_{2}=1.1432>1}\) . (a) Time series of \(S(t)\), \(x(t)\), \(F(t)\). (b) Three-dimensional phase diagrams of \(S(t)\), \(x(t)\), \(F(t)\). (c) Phase diagrams of \(x(t)\), \(F(t)\). (d) Phase diagrams of \(x(t)\), \(S(t)\).

References

  1. Novick, A, Szilard, L: Description of the chemostat. Science 112(2920), 715-716 (1950)

    Article  Google Scholar 

  2. James, TW: Continuous culture of microorganisms. Annu. Rev. Microbiol. 15(1), 27-46 (1961)

    Article  Google Scholar 

  3. Monod, J: Recherches sur la croissance des cultures bacteriennes. Hermann, Paris (1942)

    MATH  Google Scholar 

  4. Butler, GJ, Wolkowicz, GSK: A mathematical model of the chemostat with a general class of functions describing nutrient uptake. SIAM J. Appl. Math. 45(1), 138-151 (1985)

    Article  MathSciNet  MATH  Google Scholar 

  5. Hsu, SB, Hubbell, S, Waltman, P: A mathematical theory for single-nutrient competition in continuous cultures of micro-organisms. SIAM J. Appl. Math. 32(2), 366-383 (1977)

    Article  MathSciNet  MATH  Google Scholar 

  6. Levins, R: Coexistence in a variable environment. Am. Nat. 114(6), 765-783 (1979)

    Article  MathSciNet  Google Scholar 

  7. Wolkowicz, GSK, Lu, Z: Global dynamics of a mathematical model of competition in the chemostat: general response functions and differential death rates. SIAM J. Appl. Math. 52(1), 222-233 (1992)

    Article  MathSciNet  MATH  Google Scholar 

  8. Smith, HL, Waltman, P: The Theory of the Chemostat: Dynamics of Microbial Competition. Cambridge University Press, Cambridge (1995)

    Book  MATH  Google Scholar 

  9. Sun, S, Chen, L: Dynamic behaviors of Monod type chemostat model with impulsive perturbation on the nutrient concentration. J. Math. Chem. 42(4), 837-847 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  10. Luo, J, Yuan, S, Zhang, W: Competition between two microorganisms in the chemostat with general variable yields and general growth rates. Int. J. Biomath. 1(4), 463-474 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  11. Zhao, Z, Chen, L, Song, X: Extinction and permanence of chemostat model with pulsed input in a polluted environment. Commun. Nonlinear Sci. Numer. Simul. 14(4), 1737-1745 (2009)

    Article  MathSciNet  MATH  Google Scholar 

  12. Teng, Z, Gao, R, Rehim, M, Wang, K: Global behaviors of Monod type chemostat model with nutrient recycling and impulsive input. J. Math. Chem. 47(1), 276-294 (2009)

    Article  MathSciNet  MATH  Google Scholar 

  13. 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(3), 865-875 (2016)

    MathSciNet  Google Scholar 

  14. Li, Z, Wang, T, Chen, L: Periodic solution of a chemostat model with Beddington-DeAnglis uptake function and impulsive state feedback control. J. Theor. Biol. 261(1), 23-32 (2009)

    Article  MathSciNet  Google Scholar 

  15. Sun, K, Tian, Y, Chen, L, Kasperski, A: Nonlinear modelling of a synchronized chemostat with impulsive state feedback control. Math. Comput. Model. 52(1), 227-240 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  16. Zhang, T, Zhang, T, Meng, X: Stability analysis of a chemostat model with maintenance energy. Appl. Math. Lett. 68, 1-7 (2017)

    Article  MathSciNet  MATH  Google Scholar 

  17. Grover, J: Resource Competition. Chapmam & Hall, London (1997)

    Book  Google Scholar 

  18. Hansen, SR, Hubbell, SP: Single-nutrient microbial competition: qualitative agreement between experimental and theoretically forecast outcomes. Science 207(4438), 1491-1493 (1980)

    Article  Google Scholar 

  19. Tilman, D: Resource Competition and Community Structure. Princeton University Press, Princeton (1982)

    Google Scholar 

  20. Li, Z, Chen, L, Liu, Z: Periodic solution of a chemostat model with variable yield and impulsive state feedback control. Appl. Math. Model. 36(3), 1255-1266 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  21. Morales, J, de la Noue, J, Picard, G: Harvesting marine microalgae species by chitosan flocculation. Aquac. Eng. 4(4), 257-270 (1985)

    Article  Google Scholar 

  22. Yan, H, Ma, C, Sun, X, Chen, J, Wang, D: Study on flocculation of Rhodopeudomonas palustris by aluminum flocculants. Chem. Bioeng. 25(6), 53-55 (2008)

    Google Scholar 

  23. Vandamme, D, Foubert, I, Muylaert, K: Flocculation as a low-cost method for harvesting microalgae for bulk biomass production. Trends Biotechnol. 31(4), 233-239 (2013)

    Article  Google Scholar 

  24. Choi, J-H, Shin, WS, Lee, S-H, Joo, D-J, Lee, J-D, Choi, SJ, Park, LS: Application of synthetic polyamine flocculants for dye wastewater treatment. Sep. Sci. Technol. 36(13), 2945-2958 (2001)

    Article  Google Scholar 

  25. Zheng, H, Gao, Z, Yin, J, Tang, X, Ji, X, Huang, H: Harvesting of microalgae by flocculation with poly (γ-glutamic acid). Bioresour. Technol. 112, 212-220 (2012)

    Article  Google Scholar 

  26. Salim, S, Bosma, R, Vermuë, MH, Wijffels, RH: Harvesting of microalgae by bio-flocculation. J. Appl. Phycol. 23(5), 849-855 (2011)

    Article  Google Scholar 

  27. Lee, CS, Robinson, J, Chong, MF: A review on application of flocculants in wastewater treatment. Process Saf. Environ. Prot. 92(6), 489-508 (2014)

    Article  Google Scholar 

  28. Rahul, R, Jha, U, Sen, G, Mishra, S: Carboxymethyl inulin: a novel flocculant for wastewater treatment. Int. J. Biol. Macromol. 63, 1-7 (2014)

    Article  Google Scholar 

  29. Daud, Z, Nasir, N, Awang, H, et al.: Treatment of biodiesel wastewater by coagulation and flocculation using polyaluminum chloride. Aust. J. Basic Appl. Sci. 7(8), 258-262 (2013)

    Google Scholar 

  30. Sen, G, Rani, GU, Mishra, S: Microwave assisted synthesis of poly(2-hydroxyethylmethacrylate) grafted agar (Ag-g-P(HEMA)) and its application as a flocculant for wastewater treatment. Front. Chem. Sci. Eng. 7(3), 312-321 (2013)

    Article  Google Scholar 

  31. Deng, S, Bai, R, Hu, X, Luo, Q: Characteristics of a bioflocculant produced by Bacillus mucilaginosus and its use in starch wastewater treatment. Appl. Microbiol. Biotechnol. 60(5), 588-593 (2003)

    Article  Google Scholar 

  32. Wang, J-P, Chen, Y-Z, Yuan, S-J, Sheng, G-P, Yu, H-Q: Synthesis and characterization of a novel cationic chitosan-based flocculant with a high water-solubility for pulp mill wastewater treatment. Water Res. 43(20), 5267-5275 (2009)

    Article  Google Scholar 

  33. Tai, X, Ma, W, Guo, S, Yan, H: A class of dynamic delayed model describing flocculation of microorganism and its theoretical analysis. Math. Pract. Theory 13(45), 198-209 (2015)

    Google Scholar 

  34. Finn, RK, Wilson, RE: Fermentation process control, population dynamics of a continuous propagator for microorganisms. J. Agric. Food Chem. 2(2), 66-69 (1954)

    Article  Google Scholar 

  35. Bush, AW, Cool, AE: The effect of time delay and growth rate inhibition in the bacterial treatment of wastewater. J. Theor. Biol. 63(2), 385-395 (1976)

    Article  Google Scholar 

  36. Cunningham, A, Nisbet, RM: Time lag and co-operativity in the transient growth dynamics of microalgae. J. Theor. Biol. 84(2), 189-203 (1980)

    Article  Google Scholar 

  37. Ellermeyer, SF: Competition in the chemostat: global asymptotic behavior of a model with delayed response in growth. SIAM J. Appl. Math. 54(2), 456-465 (1994)

    Article  MathSciNet  MATH  Google Scholar 

  38. Freedman, HI, So, JW-H, Waltman, P: Coexistence in a model of competition in the chemostat incorporating discrete delays. SIAM J. Appl. Math. 49(3), 859-870 (1989)

    Article  MathSciNet  MATH  Google Scholar 

  39. Gopalsamy, K, Aggarwala, BD: Limit cycles in two spacies competition with time delays. J. Aust. Math. Soc. Ser. B, Appl. Math 22(2), 148-160 (1980)

    Article  MATH  Google Scholar 

  40. Macdonald, N: Time delay in simple chemostat models. Biotechnol. Bioeng. 18(6), 805-812 (1976)

    Article  Google Scholar 

  41. Ruan, S: The effect of delays on stability and persistence in plankton models. Nonlinear Anal. 24(4), 575-585 (1995)

    Article  MathSciNet  MATH  Google Scholar 

  42. Jiao, J, Yang, X, Chen, L, Cai, S: Effect of delayed response in growth on the dynamics of a chemostat model with impulsive input. Chaos Solitons Fractals 42(4), 2280-2287 (2009)

    Article  MathSciNet  MATH  Google Scholar 

  43. Meng, X, Gao, Q, Li, Z: The effects of delayed growth response on the dynamic behaviors of the Monod type chemostat model with impulsive input nutrient concentration. Nonlinear Anal., Real World Appl. 11(5), 4476-4486 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  44. Dong, Q, Ma, W, Sun, M: The asymptotic behavior of a chemostat model with Crowley-Martin type functional response and time delays. J. Math. Chem. 51(5), 1231-1248 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  45. Cao, J, Bao, J, Wang, P: Global analysis of a delayed Monod type chemostat model with impulsive input on two substrates. Adv. Differ. Equ. 2015, Article ID 310 (2015)

    Article  MathSciNet  Google Scholar 

  46. Zhang, T, Meng, X, Zhang, T, Song, Y: Global dynamics for a new high-dimensional SIR model with distributed delay. Appl. Math. Comput. 218(24), 11806-11819 (2012)

    MathSciNet  MATH  Google Scholar 

  47. Zhang, T, Meng, X, Song, Y, Zhang, T: A stage-structured predator-prey SI model with disease in the prey and impulsive effects. Math. Model. Anal. 18(4), 505-528 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  48. Braverman, E, Mamdani, R: Continuous versus pulse harvesting for population models in constant and variable environment. J. Math. Biol. 57(3), 413-434 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  49. Tian, Y, Sun, K, Chen, L, Kasperski, A: Studies on the dynamics of a continuous bioprocess with impulsive state feedback control. Chem. Eng. J. 157(2), 558-567 (2010)

    Article  Google Scholar 

  50. Guo, H, Chen, L: Periodic solution of a chemostat model with Monod growth rate and impulsive state feedback control. J. Theor. Biol. 260(4), 502-509 (2009)

    Article  MathSciNet  Google Scholar 

  51. Cheng, H, Zhang, T: A new predator-prey model with a profitless delay of digestion and impulsive perturbation on the prey. Appl. Math. Comput. 217(22), 9198-9208 (2011)

    MathSciNet  MATH  Google Scholar 

  52. Lakshmikantham, V, Bainov, DD, Simeonov, PS: Theory of Impulsive Differential Equations. World Scientific, Singapore (1989)

    Book  MATH  Google Scholar 

  53. Bainov DD, Simeonov PS: Systems with Impulse Effect: Stability, Theory, and Applications. E. Horwood Limited, Chichester (1989)

    MATH  Google Scholar 

  54. Kuang, Y: Delay Differential Equations: With Applications in Population Dynamics. Academic Press, New York (1993)

    MATH  Google Scholar 

Download references

Acknowledgements

The first author and the third author were supported by the National Natural Science Foundation of China (No. 11371230), Shandong Provincial Natural Science Foundation, China (No. ZR2015AQ001), the Project for Higher Educational Science and Technology Program of Shandong Province of China (No. J13LI05), SDUST Research Fund (2014TDJH102), and Joint Innovative Center for Safe and Effective Mining Technology and Equipment of Coal Resources, Shandong Province of China. The second author was supported by the National Natural Science Foundation of China (No. 11471034).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Wanbiao Ma.

Additional information

Competing interests

The authors declare that they have no competing interests.

Authors’ contributions

TZ and WM designed the study and carried out the analysis. TZ, WM and XM contributed to writing the paper. TZ performed numerical simulations. All authors read and approved the final manuscript.

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This article is distributed under the terms of the Creative Commons Attribution 4.0 International License (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted use, distribution, and reproduction in any medium, provided you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Zhang, T., Ma, W. & Meng, X. Global dynamics of a delayed chemostat model with harvest by impulsive flocculant input. Adv Differ Equ 2017, 115 (2017). https://doi.org/10.1186/s13662-017-1163-9

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s13662-017-1163-9

MSC

Keywords