Abstract

The aim of this paper is to study the dynamics of predator-prey interaction in a chemostat to determine whether including a discrete delay to model the time between the capture of the prey and its conversion to viable biomass can introduce oscillatory dynamics even though there is a globally asymptotically stable equilibrium when the delay is ignored. Hence, Holling type I response functions are chosen so that no oscillatory behavior is possible when there is no delay. It is proven that unlike the analogous model for competition, as the parameter modeling the delay is increased, Hopf bifurcations can occur.

1. Introduction

The chemostat, also known as a continuous stir tank reactor (CSTR) in the engineering literature, is a basic piece of laboratory apparatus used for the continuous culture of microorganisms. It has potential applications for such processes as wastewater decomposition and water purification. Some ecologists consider it a lake in a laboratory. It can be thought of as three vessels, the feed bottle that contains fresh medium with all the necessary nutrients, the growth chamber where the microorganisms interact, and the collection vessel. The fresh medium from the feed bottle is continuously added to the growth chamber. The growth chamber is well stirred and its contents are then removed to the collection vessel at a rate that maintains constant volume. For a detailed description of the importance of the chemostat and its application in biology and ecology, one can refer to [1, 2].

The following system describes a food chain in the chemostat where a predator population feeds on a prey population of microorganisms that in turn consumes a nonreproducing nutrient that is assumed to be growth limiting at low concentrations

𝑠̇𝑠(𝑡)=0𝐷𝑠(𝑡)0𝑥(𝑡)𝑓(𝑠(𝑡))𝜂,𝑦̇𝑥(𝑡)=𝑥(𝑡)(𝐷+𝑓(𝑠(𝑡)))(𝑡)𝑔(𝑥(𝑡))𝜉,̇𝑦(𝑡)=Δ𝑦(𝑡)+𝑦(𝑡)𝑔(𝑥(𝑡)).(1.1) Here 𝑠(𝑡) represents the concentration of the growth limiting nutrient, 𝑥(𝑡) the density of the prey population, and 𝑦(𝑡) the density of the predator population. Parameter 𝑠0 denotes the concentration of the growth limiting nutrient in the feed vessel, 𝐷0 the dilution rate, 𝜂(𝜉) the growth yield constant, 𝐷(Δ) the sum of the dilution rate 𝐷0 and the natural species specific death rate of the prey (predator) population, respectively. Here 𝑓(𝑠) denotes the functional response of the prey population on the nutrient and 𝑔(𝑥) denotes the functional response of the predator on the prey.

Butler et al. [3] considered the coexistence of two competing predators feeding on a single prey population growing in the chemostat. As a subsystem of their model, they studied the global stability of system (1.1) with both 𝑓(𝑠) and 𝑔(𝑥) taking the form of Holling type II. They proved that under certain conditions the interior equilibrium is globally asymptotically stable with respect to the interior of the positive cone. However, they also proved that for certain ranges of the parameters there is at least one nontrivial limit cycle and conjectured that the limit cycle is unique and would be a global attractor with respect to the noncritical orbits in the open positive octant. This conjecture was partially solved by Kuang [4]. He showed that there is a range of parameters for which a unique periodic orbit exists and roughly located the position of the limit cycle.

Bulter and Wolkowicz [5] studied predator mediated coexistence in the chemostat assuming 𝐷0=𝐷=Δ. Model (1.1) was studied as a submodel. For general monotone response functions, Bulter and Wolkowicz showed that (1.1) is uniformly persistent if the sum of the break even concentrations of substrate and prey is less than the input rate of the nutrient 𝑠0. However they showed that it is necessary to specify the form of the response functions in order to discuss the global dynamics of the model. If 𝑓(𝑠) is modelled by Holling type I or II and 𝑔(𝑥) by Holling type I, Bulter and Wolkowicz proved that (1.1) could have up to three equilibrium points and that there is a transfer of global stability from one equilibrium point to another as different parameters are varied making conditions favorable enough for a new population to survive. In this case, there are no periodic solutions. However, even if 𝑓(𝑠) is given by Holling type I, if 𝑔(𝑥) is given by Holling type II, they showed that a Hopf bifurcation can occur in (1.1), and numerical simulations indicated that the bifurcating periodic solution was asymptotically stable.

We include a time delay in (1.1) to model the time between the capture of the prey and its conversion to viable biomass. Our aim is to show that such a delay can induce nontrivial periodic solutions in a model where there is always a globally asymptotically stable equilibrium when delay is ignored, and hence no such periodic solutions are possible otherwise. For this reason we select the response functions of the simplest form; that is, we choose the Holling type I form for both 𝑓(𝑠) and 𝑔(𝑥), so that (1.2) always has a globally asymptotically stable equilibrium when the conversion process is assumed to occur instantaneously. It is interesting to note that in the analogous model of competition between two species in the chemostat, delay cannot induce oscillatory behavior for any reasonable monotone response functions (see Wolkowicz and Xia [6]).

With delay modelling the time required for the predator to process the prey after it has been captured, the model is given by

𝑠̇𝑠(𝑡)=0𝐷𝑠(𝑡)0𝑥(𝑡)𝑓(𝑠(𝑡))𝜂,𝑦̇𝑥(𝑡)=𝑥(𝑡)(𝐷+𝑓(𝑠(𝑡)))(𝑡)𝑔(𝑥(𝑡))𝜉,̇𝑦(𝑡)=Δ𝑦(𝑡)+𝑒Δ𝜏𝑦(𝑡𝜏)𝑔(𝑥(𝑡𝜏)).(1.2) For 𝑡[𝜏,0],

𝑠(0)=𝑠0int+[],(𝑥(𝑡),𝑦(𝑡))=(𝜙,𝜓)𝜏,0,int2+.(1.3) Here variables 𝑠(𝑡), 𝑥(𝑡), 𝑦(𝑡), and parameters 𝑠0, 𝐷0, 𝜂, 𝜉, 𝐷0, 𝐷, and Δ have the same interpretation as for model (1.1). Note therefore that 𝐷𝐷0, and Δ𝐷0. The additional parameter 𝜏 is a nonnegative constant modelling the time required for the conversion process. Hence, 𝑒Δ𝜏𝑦(𝑡𝜏) represents the concentration of the predator population in the growth chamber at time 𝑡 that were available at time 𝑡𝜏 to capture prey and were able to avoid death and washout during the 𝜏 units of time required to process the captured prey.

We analyze the stability of each equilibrium and prove that the coexistence equilibrium can undergo Hopf bifurcations. Numerical simulations appear to show that (1.2) can have a stable periodic solution bifurcating from the coexistence equilibrium as the delay parameter increases from zero. This periodic orbit can then disappear through a secondary Hopf bifurcation as the delay parameter increases further.

2. Scaling of the Model and Existence of Solutions

Suppose that functions 𝑓(𝑠) and 𝑔(𝑠) are of Holling type I form, that is, 𝑓(𝑠)=𝛼𝑠 (𝛼>0) and 𝑔(𝑥)=𝑘𝑥 (𝑘>0). System (1.2) reduces to

𝑠̇𝑠(𝑡)=0𝐷𝑠(𝑡)0𝛼𝑥(𝑡)𝑠(𝑡)𝜂,̇𝑥(𝑡)=𝑥(𝑡)(𝐷+𝛼𝑠(𝑡))𝑘𝑥(𝑡)𝑦(𝑡)𝜉,̇𝑦(𝑡)=Δ𝑦(𝑡)+𝑘𝑒Δ𝜏𝑦(𝑡𝜏)𝑥(𝑡𝜏).𝑡>0,(2.1) Introducing the following change of variables gives:

̆𝑡=𝐷0̆𝑡=𝑡,̆𝑠𝑠(𝑡)𝑠0̆𝑡=,̆𝑥𝑥(𝑡)𝑠0𝜂̆𝑡=,̆𝑦𝑦(𝑡)𝜉𝑠0𝜂,̆𝜏=𝐷0̆𝐷𝜏,𝐷=𝐷0,̆ΔΔ=𝐷0,̆𝑘=𝑘𝑠0𝜂𝐷0,̆𝛼=𝛼𝑠0𝐷0,̆𝑡d̆𝑠d̆𝑡=1𝑠0d𝑠(𝑡)d𝑡d𝑡d̆𝑡=1𝑠0𝐷0d𝑠(𝑡)=1d𝑡𝑠0𝐷0𝑠0𝐷𝑠(𝑡)0𝛼𝑥(𝑡)𝑠(𝑡)𝜂=1𝑠(𝑡)𝑠0𝛼𝑠0𝐷0𝑥(𝑡)𝑠0𝜂𝑠(𝑡)𝑠0̆𝑡̆𝑡̆𝑡,̆𝑡=1̆𝑠̆𝛼̆𝑥̆𝑠d̆𝑥d̆𝑡=1𝑠0𝜂d𝑥(𝑡)d𝑡d𝑡d̆𝑡=1𝑠0𝜂𝐷0d𝑥(𝑡)=1d𝑡𝑠0𝜂𝐷0𝑥(𝑡)(𝐷+𝛼𝑠(𝑡))𝑘𝑥(𝑡)𝑦(𝑡)𝜉=𝑥(𝑡)𝑠0𝜂𝐷𝐷0+𝛼𝑠0𝐷0𝑠(𝑡)𝑠0𝑘𝑠0𝜂𝐷0𝑥(𝑡)𝑠0𝜂𝑦(𝑡)𝑠0̆𝑡̆̆𝑡̆̆𝑡̆𝑡,̆𝑡𝜂𝜉=̆𝑥𝐷+̆𝛼̆𝑠𝑘̆𝑥̆𝑦d̆𝑦d̆𝑡=1𝑠0𝜂𝜉d𝑦(𝑡)d𝑡d𝑡d̆𝑡=1𝑠0𝜂𝜉𝐷0d𝑦(𝑡)=1d𝑡𝑠0𝜂𝜉𝐷0Δ𝑦(𝑡)+𝑘𝑒Δ𝜏=𝑦(𝑡𝜏)𝑥(𝑡𝜏)Δ𝑦(𝑡)𝑠0𝜂𝜉𝐷0+𝑘𝑒Δ𝜏𝑠0𝜂𝜉𝐷0=𝑦(𝑡𝜏)𝑥(𝑡𝜏)Δ𝐷0𝑦(𝑡)𝑠0+𝜂𝜉𝑘𝑠0𝜂𝐷0𝑒(Δ/𝐷0)𝐷0𝜏𝑦(𝑡𝜏)𝑠0𝜂𝜉𝑥(𝑡𝜏)𝑠0𝜂̆̆𝑡+̆=Δ̆𝑦𝑘𝑒̆Δ̆𝜏̆̆.̆𝑦𝑡̆𝜏̆𝑥𝑡̆𝜏(2.2) With this change of variables, omitting the ̆’s for convenience, system (2.1) becomes

̇𝑠(𝑡)=1𝑠(𝑡)𝛼𝑥(𝑡)𝑠(𝑡),̇𝑥(𝑡)=𝑥(𝑡)(𝐷+𝛼𝑠(𝑡))𝑘𝑦(𝑡)𝑥(𝑡),̇𝑦(𝑡)=Δ𝑦(𝑡)+𝑘𝑒Δ𝜏𝑦(𝑡𝜏)𝑥(𝑡𝜏),(2.3) where Δ1 and 𝐷1, with initial data given by (1.3). For biological significance, a point is assumed to be an equilibrium point of (2.3) only if all of its components are nonnegative.

Let 𝜏=0. Model (2.3) reduces to a special case of the model considered in [7]. If 𝐷>𝛼, the model has only one equilibrium point (1,0,0) and it is globally asymptotically stable. If 𝐷<𝛼 and 1𝐷/𝛼Δ𝐷/𝑘<0, the model has a second equilibrium point (𝐷/𝛼,(𝛼𝐷)/𝛼𝐷,0) and it is globally asymptotically stable. When 1𝐷/𝛼Δ𝐷/𝑘>0, the model has a third equilibrium point (𝑘/(𝑘+𝛼Δ),Δ/𝑘,𝛼/(𝑘+𝛼𝐷)𝐷/𝑘) and it is the global attractor. Therefore, model (2.3) has no periodic solutions when the time delay is ignored. If 𝑔(𝑥) is of Holling type II form, Butler and Wolkowicz [5] proved that a Hopf bifurcation is possible resulting in a periodic solution for a certain range of parameter values. We emphasize again here, that it is for this reason that in this paper we restrict our attention to the simplest case for both response functions, that is, Holling type I, in order to see whether delay can be responsible for periodic solutions in (1.2).

Theorem 2.1. Assuming (𝑠0,𝜙(𝜃),𝜓(𝜃))int+×([𝜏,0],int2+), then there exists a unique solution (𝑠(𝑡),𝑥(𝑡),𝑦(𝑡)) of (2.3) passing through (𝑠0,𝜙(𝜃),𝜓(𝜃)) with 𝑠(𝑡)>0, 𝑥(𝑡)>0 and 𝑦(𝑡)>0 for 𝑡[0,). The solution is bounded. In particular, given any 𝜖0>0, 𝑥(𝑡)<1+𝜖0 for all sufficiently large 𝑡.

Proof. For 𝑡[0,𝜏], one has 𝑡𝜏[𝜏,0],𝑥(𝑡𝜏)=𝜙(𝑡𝜏), and 𝑦(𝑡𝜏)=𝜓(𝑡𝜏). System (2.3) becomes ̇𝑠(𝑡)=1𝑠(𝑡)𝛼𝑥(𝑡)𝑠(𝑡),̇𝑥(𝑡)=𝑥(𝑡)(𝐷+𝛼𝑠(𝑡))𝑘𝑦(𝑡)𝑥(𝑡),̇𝑦(𝑡)=Δ𝑦(𝑡)+𝑘𝑒Δ𝜏𝜙(𝑡𝜏)𝜓(𝑡𝜏),(2.4) a system of nonautonomous ordinary differential equations with initial conditions 𝑠(0)=𝑠0, 𝑥(0)=𝜙(0), and 𝑦(0)=𝜓(0). Since the right-hand side of (2.4) is differentiable in both 𝑥 and 𝑦, by Theorems 2.3, 3.1, and Corollary 4.3 in Miller and Michel [8], there exists a unique solution defined on [0,𝜏] satisfying (2.4). By using the method of steps in Bellman and Cooke [9], it can be shown that the solution through (𝑠0,𝜙(𝜃),𝜓(𝜃)) is defined for all 𝑡0.
Now we prove 𝑠(𝑡)>0 for all 𝑡>0. From the first equation of (2.3),
̇𝑠(𝑡)=1𝑠(𝑡)𝛼𝑥(𝑡)𝑠(𝑡).(2.5) Proceed using the method of contradiction. Suppose that there exists a first 𝑡 such that 𝑠(𝑡)=0 and 𝑠(𝑡)>0 for 𝑡[0,𝑡). Then ̇𝑠(𝑡)0. But from the first equation of (2.3) 𝑡̇𝑠𝑡=1𝑠𝑡𝛼𝑥𝑠𝑡=1>0,(2.6) a contradiction.
To prove 𝑥(𝑡)>0 for 𝑡[0,), assume there is a first 𝑡>0 such that 𝑥(𝑡)=0, and 𝑥(𝑡)>0 for 𝑡[0,𝑡). Divide both sides of the second equation of (2.3) by 𝑥(𝑡) and integrate from 0 to 𝑡, to obtain
𝑥𝑡=𝜙(0)exp𝑡0(𝐷+𝛼𝑠(𝑡)𝑘𝑦(𝑡))d𝑡>0,(2.7) contradicting 𝑥(𝑡)=0.
To show that 𝑦(𝑡) is positive on [0,), suppose that there exists 𝑡>0 such that 𝑦(𝑡)=0, and 𝑦(𝑡)>0 for 𝑡[0,𝑡). Then ̇𝑦(𝑡)0. From the third equation of (2.3), we have
𝑡̇𝑦𝑡=Δ𝑦+𝑘𝑒Δ𝜏𝑦𝑡𝑥𝑡𝜏𝜏=𝑘𝑒Δ𝜏𝑦𝑡𝑥𝑡𝜏𝜏>0,(2.8) a contradiction.
To prove the boundedness of solutions, define
𝜔(𝑡)=𝑠(𝑡)+𝑥(𝑡)+𝑒Δ𝜏𝑦(𝑡+𝜏)1,for𝑡0.(2.9) It follows that ̇𝜔(𝑡)=1𝑠(𝑡)𝐷𝑥(𝑡)Δ𝑒Δ𝜏𝑦(𝑡+𝜏)1𝑠(𝑡)𝑥(𝑡)𝑒Δ𝜏𝑦(𝑡+𝜏)𝜔(𝑡),(2.10) where the first inequality holds since 𝐷1, Δ1, 𝑥(𝑡)>0 and 𝑦(𝑡+𝜏)>0. It follows that 𝑠(𝑡)+𝑥(𝑡)+𝑒Δ𝜏𝑠𝑦(𝑡+𝜏)1+0+𝑥(0)+𝑒Δ𝜏𝑒𝑦(𝜏)1𝑡1as𝑡.(2.11) Therefore, the solution (𝑠(𝑡),𝑥(𝑡),𝑦(𝑡)) is bounded, and given any 𝜖0>0, 𝑥(𝑡)<1+𝜖0 for all sufficiently large 𝑡.

3. Equilibria and Stability

Model (2.3) has three equilibrium points: 𝐸1=(1,0,0),𝐸2=(𝐷/𝛼,(𝛼𝐷)/𝛼𝐷,0), and

𝐸+=𝑠+(𝜏),𝑥+(𝜏),𝑦+=1(𝜏)1+(𝛼Δ/𝑘)𝑒Δ𝜏,Δ𝑘𝑒Δ𝜏,𝛼𝑘+𝛼Δ𝑒Δ𝜏𝐷𝑘.(3.1) We call 𝐸1 the washout equilibrium, 𝐸2 the single species equilibrium, and 𝐸+ the coexistence equilibrium. For the sake of biological significance, 𝐸+ exists (distinct from 𝐸2) if and only if its third coordinate 𝑦+(𝜏)=(𝛼𝑠+(𝜏)𝐷)/𝑘>0, that is, 𝑠+(𝜏)>𝐷/𝛼, or equivalently, 𝜏 lies between 0 and 𝜏𝑐, where

𝜏𝑐=1Δ𝑘lnΔ1𝐷1𝛼.(3.2) Note that if (𝑘/Δ)(1/𝐷1/𝛼)1, the equilibrium 𝐸+ does not exist for any 𝜏 (0), and if (𝑘/Δ)(1/𝐷1/𝛼)=1, then 𝐸+=𝐸2.

The linearization of (2.3) about an equilibrium (𝑠,𝑥,𝑦) is given by

̇𝑧1(𝑡)̇𝑧2(𝑡)̇𝑧3=𝑧(𝑡)1𝛼𝑥𝛼𝑠0𝛼𝑥𝐷+𝛼𝑠𝑘𝑦𝑘𝑥00Δ1𝑧(𝑡)2𝑧(𝑡)3+(𝑡)0000000𝑘𝑒Δ𝜏𝑦𝑘𝑒Δ𝜏𝑥𝑧1𝑧(𝑡𝜏)2𝑧(𝑡𝜏)3.(𝑡𝜏)(3.3) The associated characteristic equation is given by

det1𝛼𝑥𝜆𝛼𝑠0𝛼𝑥𝐷+𝛼𝑠𝑘𝑦𝜆𝑘𝑥0𝑘𝑒Δ𝜏𝜆𝜏𝑦Δ+𝑘𝑒Δ𝜏𝜆𝜏𝑥𝜆=0.(3.4) Direct calculation of the left-hand side of (3.4) gives

Δ+𝑘𝑒(Δ+𝜆)𝜏𝑥𝜆(1𝛼𝑥𝜆)(𝐷+𝛼𝑠𝑘𝑦𝜆)+𝛼2𝑠𝑥+𝑘𝑥𝑘𝑒(Δ+𝜆)𝜏𝑦((1𝛼𝑥𝜆)=(Δ𝜆)1+𝛼𝑥+𝜆)(𝐷𝛼𝑠+𝑘𝑦+𝜆)+𝛼2𝑠𝑥+𝑒(Δ+𝜆)𝜏×𝑘𝑥𝑘𝑦(1𝛼𝑥𝜆)+(1+𝛼𝑥+𝜆)(𝐷𝛼𝑠+𝑘𝑦+𝜆)+𝛼2=𝑠𝑥(Δ𝜆)(1+𝛼𝑥+𝜆)(𝐷𝛼𝑠+𝑘𝑦+𝜆)+𝛼2𝑠𝑥+𝑒(Δ+𝜆)𝜏𝑘𝑥(1+𝛼𝑥+𝜆)(𝐷𝛼𝑠+𝜆)+𝛼2=𝑠𝑥(Δ𝜆){(𝜆+1)(𝜆+𝐷+𝑘𝑦)+𝛼𝑥(𝜆+𝐷+𝑘𝑦)𝛼𝑠(𝜆+1)}+𝑒(Δ+𝜆)𝜏𝑘𝑥{(𝜆+1)(𝜆+𝐷)+𝛼𝑥(𝜆+𝐷)𝛼𝑠(𝜆+1)}.(3.5) For convenience, define 𝑃(𝜆)=(Δ𝜆){(𝜆+1)(𝜆+𝐷+𝑘𝑦)+𝛼𝑥(𝜆+𝐷+𝑘𝑦)𝛼𝑠(𝜆+1)}+𝑒(Δ+𝜆)𝜏𝑘𝑥{(𝜆+1)(𝜆+𝐷)+𝛼𝑥(𝜆+𝐷)𝛼𝑠(𝜆+1)}.(3.6)

Theorem 3.1. Equilibrium 𝐸1 is stable if 𝛼<𝐷 and unstable if 𝛼>𝐷.

Proof. Evaluating the characteristic equation at 𝐸1 gives ||𝑃(𝜆)𝐸1=(Δ+𝜆)(𝜆+1)(𝜆+𝐷𝛼)=0.(3.7) The eigenvalues 1 and Δ are both negative. The third eigenvalue is 𝐷+𝛼. Therefore the equilibrium 𝐸1 is stable if 𝛼<𝐷 and unstable if 𝛼>𝐷.

Remark 3.2. If 𝛼<𝐷, then there is only one equilibrium, 𝐸1. If 𝛼>𝐷, equilibrium 𝐸2 also exists.

Lemma 3.3. Assume 𝛼>𝐷. The characteristic equation evaluated at 𝐸2 has two negative eigenvalues, and the remaining eigenvalues are solutions of (𝜆+Δ)𝑒(𝜆+Δ)𝜏1=𝑘𝐷1𝛼.(3.8) In addition, the characteristic equation evaluated at 𝐸2 has zero as an eigenvalue if and only if 𝜏=𝜏𝑐.

Proof. Assume 𝛼>𝐷. Equilibrium 𝐸2 exists. Consider the characteristic equation at 𝐸2. Since (𝛼𝐷)/𝛼𝐷=(1𝑠)/𝛼𝑠 at 𝐸2, ||𝑃(𝜆)𝐸2×={(𝜆+1)(𝜆+𝐷)+𝛼𝑥(𝜆+𝐷)𝛼𝑠(𝜆+1)}𝜆Δ+𝑒(Δ+𝜆)𝜏=𝑘𝑥(𝜆+1)(𝜆+𝐷)+1𝑠𝑠×(𝜆+𝐷)𝐷(𝜆+1)𝜆Δ+𝑒(Δ+𝜆)𝜏𝑘𝛼𝐷=𝛼𝐷𝜆(𝜆+1)(𝜆+𝐷)+𝜆+𝐷𝑠𝜆Δ+𝑘𝛼𝐷𝑒𝛼𝐷(Δ+𝜆)𝜏𝜆=2+𝛼𝐷𝜆+𝛼𝐷𝜆+Δ𝑘𝛼𝐷𝑒𝛼𝐷(Δ+𝜆)𝜏=𝑒(Δ+𝜆)𝜏𝜆𝜆1𝜆𝜆2(𝜆+Δ)𝑒(Δ+𝜆)𝜏1𝑘𝐷1𝛼=0,(3.9) where 𝜆1+𝜆2=𝛼/𝐷 and 𝜆1𝜆2=𝛼𝐷>0. Therefore, 𝜆1 and 𝜆2 have negative real parts. The rest of the eigenvalues are roots of (3.8).
Assuming that 𝜆=0 is a root of (3.8), we have
Δ𝑒Δ𝜏1=𝑘𝐷1𝛼.(3.10) Solving for 𝜏 gives 1𝜏=Δ𝑘lnΔ1𝐷1𝛼=𝜏𝑐.(3.11)

Theorem 3.4. Assume that 𝐷1, Δ1, 𝑘>0, 𝛼>0, and (𝑘/Δ)(1/𝐷1/𝛼)1 so that 𝜏𝑐0. Equilibrium 𝐸2 is locally asymptotically stable if 𝜏>𝜏𝑐 and unstable if 𝜏<𝜏𝑐. If 𝐷=1, then equilibrium 𝐸2 is globally asymptotically stable for 𝜏>(1/Δ)ln(𝑘/Δ).

Proof. Assume that 𝜏>𝜏𝑐. Assumptions 𝑘>0, Δ1, and (𝑘/Δ)(1/𝐷1/𝛼)1 imply 1/𝐷>1/𝛼, or equivalently 𝛼>𝐷. By Lemma 3.3, to prove that equilibrium 𝐸2 is locally asymptotically stable, one only needs to show that (3.8) admits no root with nonnegative real part.
Consider the real roots of (3.8) first. Note that 1/𝐷>1/𝛼. Equation (3.8) has no solution for 𝜆Δ. Otherwise the left-hand side would be less than zero, but the right-hand side would be greater than zero. Assume 𝜆>Δ. The left-hand side of (3.8) is a monotone increasing function in both 𝜆 and 𝜏, takes value 0 at 𝜆=Δ, and goes to positive infinity as 𝜆+ or 𝜏+. By Lemma 3.3, when 𝜏=𝜏𝑐, then 𝜆=0 is a solution of (3.8). Thus for 𝜏>𝜏𝑐, any real root 𝜆 of (3.8) must satisfy Δ<𝜆<0.
For any 𝜏=̃𝜏<𝜏𝑐, we have (𝜆+Δ)𝑒(𝜆+Δ)𝜏|𝜏=̃𝜏,𝜆=0<𝑘(1/𝐷1/𝛼) and lim𝜆+(𝜆+Δ)𝑒(𝜆+Δ)̃𝜏=+. Therefore there exists at least one ̃𝜆=𝜆>0 such that ̃(̃𝜏,𝜆) is a solution of (3.8). Equilibrium 𝐸2 is unstable if 𝜏<𝜏𝑐.
In what follows, we prove that if 𝜏>𝜏𝑐 all complex eigenvalues of (3.8) have negative real parts. Suppose that 𝜆+Δ=𝛾+𝑖𝛽(𝛽>0) is a solution of (3.8). Using the Euler formula, we have
1𝛾cos(𝛽𝜏)𝛽sin(𝛽𝜏)+𝑖(𝛾sin(𝛽𝜏)+𝛽cos(𝛽𝜏))=𝑘𝐷1𝛼𝑒𝛾𝜏.(3.12) Equating the real parts and imaginary parts of the equation, we have 1𝛾cos(𝛽𝜏)𝛽sin(𝛽𝜏)=𝑘𝐷1𝛼𝑒𝛾𝜏𝛾sin(𝛽𝜏)+𝛽cos(𝛽𝜏)=0.(3.13) Squaring both equations, adding, and taking the square root on both sides give 𝛾2+𝛽2𝑒𝛾𝜏1=𝑘𝐷1𝛼.(3.14) The left-hand side of (3.14) is monotonically increasing in 𝛾, 𝛽, and 𝜏 provided that 𝛾>0. Since (3.14) has solution 𝛾=Δ, 𝛽=0 at 𝜏=𝜏𝑐, any roots of (3.14) must satisfy 𝛾<Δ since 𝜏>𝜏𝑐. Hence Re{𝜆}=𝛾Δ<0. Therefore (3.8) has no complex eigenvalue with nonnegative real part and so 𝐸2 is locally asymptotically stable for 𝜏>𝜏𝑐.
Assume that 𝐷=1. Now we prove that 𝐸2 is globally asymptotically stable when 𝜏>(1/Δ)ln(𝑘/Δ), or equivalently 𝑘𝑒Δ𝜏<Δ. In this case, choose 𝜖0>0 small enough such that 𝑘𝑒Δ𝜏(1+𝜖0)<Δ. By Theorem 2.1, for such 𝜖0, there exists a 𝑇>0 so that 0<𝑥(𝑡)<1+𝜖0 for 𝑡>𝑇. Hence, for 𝑡>𝑇+𝜏, 𝑘𝑒Δ𝜏𝑥(𝑡𝜏)<Δ. In Example 5.1 of Kuang ([10, page 32]), choose 𝜌(𝑡)=𝜏, 𝑎(𝑡)=Δ, 𝑏(𝑡)=𝑘𝑒Δ𝜏𝑥(𝑡𝜏), and 𝛼=Δ/2. We obtain (𝑘𝑒Δ𝜏(1+𝜖0))2<Δ2=4(Δ𝛼)𝛼. Therefore 𝑦(𝑡)0 as 𝑡. Let 𝑧(𝑡)=𝑠(𝑡)+𝑥(𝑡). Noting 𝐷=1, from (2.3), we have ̇𝑧(𝑡)=1𝑧(𝑡)𝑘𝑥(𝑡)𝑦(𝑡). Multiply by the integrating factor 𝑒𝑡, (𝑧(𝑡)𝑒𝑡)=𝑒𝑡(1𝑘𝑥(𝑡)𝑦(𝑡)). Integrating both sides from 0 to 𝑡 gives
𝑧(𝑡)=𝑒𝑡𝑧(0)+𝑒𝑡𝑒𝑡1𝑒𝑡𝑡0𝑒𝑠𝑘𝑥(𝑠)𝑦(𝑠)d𝑠=1+𝑒𝑡(𝑧(0)1)𝑒𝑡𝑡0𝑒𝑠𝑘𝑥(𝑠)𝑦(𝑠)d𝑠.(3.15) If lim𝑡𝑡0𝑒𝑠𝑘𝑥(𝑠)𝑦(𝑠)d𝑠<, then lim𝑡𝑒𝑡𝑡0𝑒𝑠𝑘𝑥(𝑠)𝑦(𝑠)d𝑠=0. Therefore lim𝑡𝑧(𝑡)=1. If lim𝑡𝑡0𝑒𝑠𝑘𝑥(𝑠)𝑦(𝑠)d𝑠=, by L'Hôspital's rule, lim𝑡+𝑒𝑡𝑡0𝑒𝑠𝑘𝑥(𝑠)𝑦(𝑠)d𝑠=lim𝑡+𝑡0𝑒𝑠𝑘𝑥(𝑠)𝑦(𝑠)d𝑠𝑒𝑡=lim𝑡+𝑒𝑡𝑘𝑥(𝑡)𝑦(𝑡)𝑒𝑡=lim𝑡+𝑘𝑥(𝑡)𝑦(𝑡)=0,(3.16) since 𝑥(𝑡) is bounded and lim𝑡𝑦(𝑡)=0. It again follows that limt𝑧(𝑡)=1. Hence lim𝑡𝑠(𝑡)+𝑥(𝑡)=1.(3.17)
We show that lim𝑡𝑠(𝑡)=1/𝛼 and lim𝑡𝑥(𝑡)=(𝛼1)/𝛼. First assume that the limits exist, that is, lim𝑡𝑠(𝑡)=𝑠 and lim𝑡𝑥(𝑡)=𝑥. From (2.3), we know that ̇𝑠(𝑡) and ̇𝑥(𝑡) are uniformly continuous since 𝑠(𝑡), 𝑥(𝑡), and 𝑦(𝑡) are bounded. By Theorem A.3, it follows that lim𝑡̇𝑠(𝑡)=0 and lim𝑡̇𝑥(𝑡)=0. Note that lim𝑡𝑦(𝑡)=0. Letting 𝑡 in (2.3) gives
1𝑠𝛼𝑥𝑠=0,𝑥1+𝛼𝑠=0.(3.18) Either (𝑠,𝑥)=(1,0) or (𝑠,𝑥)=(1/𝛼,(𝛼1)/𝛼). Assume that (𝑠,𝑥)=(1,0), that is, lim𝑡𝑠(𝑡)=1 and lim𝑡𝑥(𝑡)=0. Note that 𝛼>𝐷. There exists 𝜖>0 such that 𝛼𝐷(𝛼+𝑘)𝜖>0. For such 𝜖, there exists a sufficiently large 𝑡 so that 𝑠(𝑡)>1𝜖 and 0<𝑦(𝑡)<𝜖. Recalling that 𝑥(𝑡)>0, by (2.3) ̇𝑥(𝑡)>𝑥(𝑡)(𝐷+𝛼(1𝜖)𝑘𝜖)=𝑥(𝑡)(𝛼𝐷𝛼𝜖𝑘𝜖)>0,(3.19) for all sufficiently large 𝑡. Therefore it is impossible for 𝑥(𝑡) to approach 0 from above giving a contradiction. Therefore, we must have (𝑠,𝑥)=(1/𝛼,(𝛼1)/𝛼).
Now suppose that the limits do not exist. In particular if 𝑥(𝑡) does not converge, then let 𝑥=limsup𝑡𝑥(𝑡) and 𝑥=liminf𝑡𝑥(𝑡). By Lemma A.2 in the appendix, there exists {𝑡𝑚} and {𝑠𝑚} such that
lim𝑚𝑥𝑡𝑚=𝑥,lim𝑚𝑡̇𝑥𝑚=0,lim𝑚𝑥𝑠𝑚=𝑥lim𝑚𝑠̇𝑥𝑚=0.(3.20) From (2.3), 𝑥𝑡𝑚𝑡𝐷+𝛼𝑠𝑚𝑡+𝑘𝑦𝑚=0.(3.21) Noting that 𝑥(𝑡𝑚)>0, we have 𝑠(𝑡𝑚)=(1𝑘𝑦(𝑡𝑚))/𝛼. Since lim𝑡𝑦(𝑡)=0, lim𝑡𝑠(𝑡𝑚)=1/𝛼. By (3.17), lim𝑡𝑥(𝑡𝑚)=lim𝑡(𝑥(𝑡𝑚)+𝑠(𝑡𝑚))𝑠(𝑡𝑚)=11/𝛼=(𝛼1)/𝛼. Therefore 𝑥=(𝛼1)/𝛼. Similarly we can show that 𝑥=(𝛼1)/𝛼. This implies that lim𝑡𝑥(𝑡)=(𝛼1)/𝛼, a contradiction.
Since 𝑠(𝑡)+𝑥(𝑡) converges and 𝑥(𝑡) converges, then 𝑠(𝑡) must also converge. Hence lim𝑡𝑠(𝑡)=1/𝛼 and lim𝑡𝑥(𝑡)=(𝛼1)/𝛼. It follows that 𝐸2 is globally asymptotically stable.

4. Hopf Bifurcations at 𝐸+ Assuming 𝐷=Δ=1

Now consider the stability of 𝐸+. The characteristic equation at 𝐸+ is

||𝑃(𝜆)𝐸+=(Δ𝜆)1+𝛼𝑥+(𝜏)+𝜆𝐷𝛼𝑠+(𝜏)+𝑘𝑦+(𝜏)+𝜆+𝛼2𝑠+(𝜏)𝑥+(𝜏)+𝑒(Δ+𝜆)𝜏𝑘𝑥+(𝜏)(𝜆+1)(𝜆+𝐷)+𝛼𝑥+(𝜏)(𝜆+𝐷)𝛼𝑠+(𝜏)(𝜆+1)=(Δ𝜆)1+𝛼𝑥+(𝜏)+𝜆𝜆+𝛼2𝑠+(𝜏)𝑥+(𝜏)+𝑒(Δ+𝜆)𝜏𝑘𝑥+(𝜏)(𝜆+1)(𝜆+𝐷)+𝛼𝑥+(𝜏)(𝜆+𝐷)𝛼𝑠+1(𝜏)(𝜆+1)=(Δ𝜆)𝑠+(𝜏)+𝜆𝜆+𝛼1𝑠+(𝜏)+𝑒𝜆𝜏Δ(𝜆+1)(𝜆+𝐷)+1𝑠+(𝜏)𝑠+(𝜏)(𝜆+𝐷)𝛼𝑠+𝜆(𝜏)(𝜆+1)=(Δ𝜆)2+𝜆𝑠+(𝜏)+𝛼1𝑠+(𝜏)+Δ𝑒𝜆𝜏1𝜆+𝑠+(𝜏)(𝜆+𝐷)𝛼𝑠+(𝜏)(𝜆+1)=0.(4.1) By assumption Δ=𝐷=1, and so

||𝑃(𝜆)𝐸+𝜆=(𝜆+1)2+𝜆𝑠+(𝜏)+𝛼1𝑠+(𝜏)+𝑒𝜆𝜏𝜆+𝛼𝑠+1(𝜏)𝑠+𝜆(𝜏)=(𝜆+1)2+𝑝(𝜏)𝜆+𝛽(𝜏)+𝑒𝜆𝜏(𝑞𝜆+𝑐(𝜏))=0,(4.2) where

1𝑝(𝜏)=𝑠+(𝜏),𝛽(𝜏)=𝛼1𝑠+(𝜏),𝑞=1,𝑐(𝜏)=𝛼𝑠+1(𝜏)𝑠+.(𝜏)(4.3) The characteristic equation at 𝐸+ has one eigenvalue equal to 1 and the others are given by solutions of the equation

𝜆2+𝑝(𝜏)𝜆+𝛽(𝜏)+𝑒𝜆𝜏(𝑞𝜆+𝑐(𝜏))=0.(4.4)

Lemma 4.1. Assuming 𝑘>0, 𝛼>0, and 𝑘(11/𝛼)1 so that 𝜏𝑐=ln(𝑘(11/𝛼))0, then 𝐸+ has no zero eigenvalue for 𝜏(0,𝜏𝑐).

Proof. Assume that 𝜏(0,𝜏𝑐). By the method of contradiction, suppose that there exists a zero root of (4.4). Therefore 1𝛽(𝜏)+𝑐(𝜏)=𝛼𝑠+(𝜏)=0.(4.5) Noting that 𝜏𝑐>0 if and only if 𝑘(11/𝛼)>1, for any 0<𝜏<𝜏𝑐, 1𝛼𝑠+𝛼(𝜏)=𝛼1𝑘𝑒𝜏1>𝛼1𝛼1𝛼=0,(4.6) a contradiction.

Lemma 4.2. Assume 𝑘>0, 𝛼>0, 𝑘(11/𝛼)>1. Equilibrium 𝐸+ is asymptotically stable when 𝜏=0.

Proof. For 𝜏=0, (4.4) reduces to 𝜆2+𝑝(0)𝜆+𝛽(0)+(𝑞𝜆+𝑐(0))=𝜆2+1𝑠+1(0)1𝜆+𝛼𝑠+(0).(4.7) Both coefficients are positive, since 1𝑠+𝛼(0)1=𝑘1>0,𝛼𝑠+𝛼(0)=𝛼1𝑘1=𝛼1𝛼1𝑘>0,(4.8) and 𝑘(11/𝛼)>1 implies 11/𝛼>1/𝑘. Therefore, all the roots of the characteristic equation have negative real parts.

Lemma 4.3. As 𝜏 is increased from 0, a root of (4.4) with positive real part can only appear if a root with negative real part crosses the imaginary axis.

Proof. Taking 𝑛=2 and 𝑔(𝜆,𝜏)=𝑝(𝜏)𝜆+(𝑞𝜆+𝑐(𝜏))𝑒𝜆𝜏+𝛽(𝜏) in Kuang [10, Theorem 1.4, page 66] gives limsupRe𝜆>0,|𝜆|||𝜆2𝑔||(𝜆,𝜏)=0<1.(4.9) Therefore, no root of (4.4) with positive real part can enter from infinity as 𝜏 increases from 0. Hence roots with positive real part can only appear by crossing the imaginary axis.
For 𝜏0, assuming 𝜆=𝑖𝜔 (𝜔>0) is a root of 𝑃(𝜆)|𝐸+=0,
𝜔2+𝑖𝑝(𝜏)𝜔+𝛽(𝜏)+𝑒𝑖𝜔𝜏(𝑖𝑞𝜔+𝑐(𝜏))=0.(4.10) Substituting 𝑒𝑖𝜃=cos𝜃+𝑖sin𝜃 into (4.10) gives 𝜔2+𝛽(𝜏)+𝑞𝜔sin(𝜔𝜏)+𝑐(𝜏)cos(𝜔𝜏)+𝑖(𝑝(𝜏)𝜔+𝑞𝜔cos(𝜔𝜏)𝑐(𝜏)sin(𝜔𝜏))=0.(4.11) Separating the real and imaginary parts, we obtain 𝑐(𝜏)cos(𝜔𝜏)+𝑞𝜔sin(𝜔𝜏)=𝜔2𝛽(𝜏),𝑐(𝜏)sin(𝜔𝜏)𝑞𝜔cos(𝜔𝜏)=𝑝(𝜏)𝜔.(4.12) Solving for cos(𝜔𝜏) and sin(𝜔𝜏) gives 𝜔sin(𝜔𝜏)=𝑐(𝜏)𝑝(𝜏)𝜔+𝑞𝜔2𝛽(𝜏)𝑐(𝜏)2+𝑞2𝜔2,𝜔cos(𝜔𝜏)=𝑐(𝜏)2𝛽(𝜏)𝑞𝑝(𝜏)𝜔2𝑐(𝜏)2+𝑞2𝜔2.(4.13) Noting sin2(𝜔𝜏)+cos2(𝜔𝜏)=1, squaring both sides of equations (4.13), adding, and rearranging gives 𝜔4+𝑝2(𝜏)𝑞2𝜔2𝛽(𝜏)2+𝛽2(𝜏)𝑐2(𝜏)=0.(4.14) Solving for 𝜔, we obtain two roots 𝜔1(𝜏) and 𝜔2(𝜏): 𝜔11(𝜏)=2𝑞2𝑝2(𝜏)+2𝛽(𝜏)+𝑞2𝑝2(𝜏)+2𝛽(𝜏)24(𝛽2(𝜏)𝑐2(𝜏))1/2=1𝑠+(𝜏)21𝑠+(𝜏)2𝛼𝑠2+(𝜏)𝑠+(+𝜏)1𝑠2+(𝜏)12+4𝛼𝑠2+𝑠(𝜏)2+(𝜏)11𝑠+(𝜏)+4𝑠2+(𝜏)𝛼𝑠2+(𝜏)121/2𝜔21(𝜏)=2𝑞2𝑝2(𝜏)+2𝛽(𝜏)𝑞2𝑝2(𝜏)+2𝛽(𝜏)2𝛽42(𝜏)𝑐2(𝜏)1/2=1𝑠+(𝜏)21𝑠+(𝜏)2𝛼𝑠2+(𝜏)𝑠+(𝜏)1𝑠2+(𝜏)12+4𝛼𝑠2+𝑠(𝜏)2+(𝜏)11𝑠+(𝜏)+4𝑠2+(𝜏)𝛼𝑠2+(𝜏)121/2.(4.15)
Define conditions (𝐻1) and (𝐻2) as follows:
𝐻1𝑞2𝑝2(𝜏)+2𝛽(𝜏)>0,𝛽2(𝜏)𝑐2𝑞(𝜏)>0,2𝑝2(𝜏)+2𝛽(𝜏)2𝛽42(𝜏)𝑐2(𝜏)0,(4.16)𝐻2𝛽2(𝜏)𝑐2(𝜏)<0,or𝛽2(𝜏)𝑐2(𝜏)=0and𝑞2𝑝2(𝜏)+2𝛽(𝜏)>0.(4.17)

Lemma 4.4. If (𝐻1) holds for all 𝜏 in some interval 𝐼, then (4.14) has two positive roots 𝜔1(𝜏)𝜔2(𝜏) for all 𝜏𝐼 with 𝜔1(𝜏)>𝜔2(𝜏) when all the inequalities in (𝐻1) are strict. If (𝐻2) holds for all 𝜏 in some interval 𝐼, then (4.14) has only one positive root, 𝜔1(𝜏) for all 𝜏𝐼. If no interval exists where either (𝐻1) or (𝐻2) holds, then there are no positive real roots of (4.14).

Define the interval

𝑘𝐽=ln𝛼11/4+𝑘1/16+1/(2𝛼)1,ln4𝛼1𝛼.(4.18) When the end points of 𝐽 are real and 𝐽, define

𝐼1=0,𝜏𝑐𝐽.(4.19) We prove that (𝐻1) holds for any 𝜏𝐼1.

From 𝐷=Δ=1,

𝜏𝑐=1Δ𝑘lnΔ1𝐷1𝛼=ln𝑘(𝛼1)𝛼.(4.20) If 𝛼>1, then 𝛼>4𝛼. It follows that

𝜏𝑐𝑘>ln4𝛼1𝛼.(4.21) Therefore,

𝐼1=𝑘max0,ln𝛼11/4+𝑘1/16+1/(2𝛼)1,ln4𝛼1𝛼.(4.22)

Theorem 4.5. Assume 𝛼>(7+35)/2 and 𝑘>𝛼/(4𝛼1), then 𝐼1 is not empty, and for any 𝜏𝐼1, but 𝜏ln((𝑘/𝛼)(1/(1/4+1/16+1/(2𝛼))1)), condition (𝐻1) holds and 𝜔1(𝜏)>𝜔2(𝜏)>0. If 𝜏=ln((𝑘/𝛼)(1/(1/4+1/16+1/(2𝛼))1))𝐼1, then 𝜔1(𝜏)>𝜔2(𝜏)=0.

Proof. For any 𝛼>(7+35)/2, we have 11/4𝛼>0, and therefore 14𝛼+1252<147+35+1/2252=0.(4.23) Hence, 14𝛼1421+1162𝛼2=1𝛼124𝛼1=2𝛼124𝛼14𝛼32+14𝛼=124𝛼114𝛼14𝛼2+14𝛼=1124𝛼114𝛼14𝛼+12254=124𝛼114𝛼14𝛼+125214𝛼+12+52<0.(4.24) Therefore, 1/4𝛼1/4<1/16+1/(2𝛼). Since 1/4+1/16+1/(2𝛼)<1/4+1/16+1/(7+35)<1, it follows that 14𝛼<14+1+1162𝛼<1.(4.25) Hence, 𝑘ln𝛼11/4+𝑘1/16+1/(2𝛼)1<ln4𝛼1𝛼.(4.26) From 𝑘>𝛼/(4𝛼1), we have ln(𝑘(4𝛼1)/𝛼)>0. Therefore, 𝑘max0,ln𝛼11/4+𝑘1/16+1/(2𝛼)1<ln4𝛼1𝛼,(4.27) and so 𝐼1 is not empty. Noting 𝑠+(𝜏)=1/(1+(𝛼Δ/𝑘)𝑒Δ𝜏) and Δ=1, for any 𝜏𝐼1, but 𝜏ln((𝑘/𝛼)(1/(1/4+1/16+1/(2𝛼))1)), we have 𝑠+(𝜏)[1/4𝛼,1/4+1/16+1/(2𝛼)).
In what follows, we intend to show that for any such 𝜏, condition (𝐻1) holds. From (4.3),
𝑞2𝑝2(𝜏)+2𝛽(𝜏)=(1)21𝑠2+(𝜏)+2𝛼1𝑠+=(𝜏)1𝑠+(𝜏)2𝛼𝑠2+𝑠(𝜏)+(𝜏)2𝑠+(𝜏)12𝛼=2𝛼1𝑠+(𝜏)2𝛼𝑠2+𝑠(𝜏)+1(𝜏)4𝛼2116𝛼21.2𝛼(4.28) Since 𝑠+(𝜏)<1, to show that the first inequality in (𝐻1) holds, it suffices to show that the factor on the right-hand side of the above expression is positive. Since 𝛼>(7+35)/2, 1/𝛼1/(4𝛼)=(1/𝛼)(11/(4𝛼))>0, and 116𝛼2+12𝛼21𝛼14𝛼2=116𝛼2+112𝛼𝛼+12𝛼𝛼116𝛼2=112𝛼𝛼1<0.(4.29) Since 1/𝛼<1/4𝛼 for 𝛼>(7+35)/2, 1+4𝛼116𝛼2+1<12𝛼𝛼<14𝛼.(4.30) For any 𝑠+(𝜏)>1/4𝛼, 𝑠+1(𝜏)14𝛼4𝛼1>4𝛼116𝛼2+1.2𝛼(4.31) Hence, 𝑠+1(𝜏)4𝛼2116𝛼2+12𝛼.(4.32) Next consider the second inequality in (𝐻1). For 𝛼>(7+35)/2, since 1/4𝛼>1/𝛼, 𝑠+(𝜏)1/4𝛼>1/𝛼. Therefore, 𝛼𝑠+(𝜏)>1. For 𝑠+(𝜏)[1/4𝛼,1/4+1/16+1/(2𝛼))𝛽2(𝜏)𝑐2=(𝜏)=(𝛽(𝜏)𝑐(𝜏))(𝛽(𝜏)+𝑐(𝜏))𝛼2𝛼𝑠+1(𝜏)+𝑠+(1𝜏)𝛼𝑠+(𝜏)=2𝛼𝑠2+𝑠(𝜏)2+𝑠(𝜏)+(𝜏)212𝛼𝛼𝑠+(𝜏)1=2𝛼𝑠2+𝑠(𝜏)+1(𝜏)4211162𝛼𝛼𝑠+(𝜏)1>0.(4.33) Finally, 𝑞2𝑝2(𝜏)+2𝛽(𝜏)2𝛽42(𝜏)𝑐2=𝑞(𝜏)2𝑝2𝑞(𝜏)2𝑝2(𝜏)+4𝛽(𝜏)+4𝑐2=1(𝜏)1𝑠2+(1𝜏)1𝑠2+(𝜏)+4𝛼1𝑠+(𝜏)+4𝛼𝑠+1(𝜏)𝑠+(𝜏)2=11𝑠2+(𝜏)21+4𝛼1𝑠2+(𝜏)1𝑠+(𝜏)+4𝑠+(𝜏)2𝛼248𝛼+𝑠2+(𝜏)=4𝑠+(𝜏)2𝛼21+4𝛼1𝑠2+(𝜏)1𝑠+1(𝜏)8𝛼+1𝑠2+(𝜏)2+4𝑠2+(𝜏)=4𝑠+(𝜏)2𝛼21+4𝛼1𝑠2+(𝜏)1𝑠++1(𝜏)21+𝑠2+(𝜏)2=𝛼𝛼1𝛼𝛼2,(4.34) where 𝛼1=211/𝑠2+(𝜏)1𝑠++(𝜏)1/𝑠2+(𝜏)+2𝑠+(𝜏)+11/𝑠+(𝜏)122𝑠2+,𝛼(𝜏)2=211/𝑠2+(𝜏)1𝑠+(𝜏)1/𝑠2+(𝜏)+2𝑠+(𝜏)+11/𝑠+(𝜏)122𝑠2+.(𝜏)(4.35) Since 𝑠+(𝜏)<1, 121𝑠2+(𝜏)1𝑠+(𝜏)=𝑠+(𝜏)+1+1𝑠+(𝜏)𝑠2+1(𝜏)>0,21𝑠2+(𝜏)1𝑠+(𝜏)2>121𝑠2+(𝜏)1𝑠+(𝜏)2𝑠2+1(𝜏)1+𝑠2+(𝜏)2=1𝑠2+(𝜏)+2𝑠+1(𝜏)+1𝑠+(𝜏)12>0.(4.36) It follows that 0<𝛼2<𝛼1. Again noting that 𝑠+(𝜏)<1, 𝛼1<211/𝑠2+(𝜏)1𝑠++(𝜏)1/𝑠2+(𝜏)+2/𝑠+(𝜏)+11/𝑠+(𝜏)122𝑠2+=(𝜏)211/𝑠2+(𝜏)1𝑠++(𝜏)1/𝑠+(𝜏)+11/𝑠+(𝜏)12𝑠2+=(𝜏)21𝑠+(𝜏)1/𝑠2+(𝜏)+1/𝑠+(𝜏)+1/𝑠2+(𝜏)12𝑠2+=𝑠(𝜏)+(𝜏)+2/𝑠2+(𝜏)1/𝑠+(𝜏)2𝑠2+=1(𝜏)22𝑠4+1(𝜏)𝑠3+1(𝜏)𝑠+<1(𝜏)22𝑠4+=1(𝜏)𝑠4+.(𝜏)(4.37) Hence, for any 𝑠+(𝜏)>1/4𝛼, we have 𝛼>1/𝑠4+(𝜏)>𝛼1>𝛼2. This leads to 𝑞2𝑝2(𝜏)+2𝛽(𝜏)2𝛽42(𝜏)𝑐2=(𝜏)𝛼𝛼1𝛼𝛼2>0.(4.38) Therefore, (𝐻1) holds for any 𝜏𝐼1. By Lemma 4.4, both 𝜔1(𝜏)>0 and 𝜔2(𝜏)>0.
If 𝜏=ln((𝑘/𝛼)(1/(1/4+1/16+1/(2𝛼))1))𝐼1, we have 𝑠+(𝜏)=1/4+1/16+1/(2𝛼). Noting (4.25), we obtain
𝑞2𝑝2(𝜏)+2𝛽(𝜏)>0,𝛽2(𝜏)𝑐2𝑞(𝜏)=0,2𝑝2(𝜏)+2𝛽(𝜏)2𝛽42(𝜏)𝑐2(𝜏)>0.(4.39) By (4.15), it follows that 𝜔1(𝜏)>0 and 𝜔2(𝜏)=0.

Now we define interval 𝐼2 and prove that (𝐻2) holds on 𝐼2

𝐼2=0,𝜏𝑐𝑘,ln𝛼11/4+1/16+1/(2𝛼)1.(4.40) In the following theorem, we consider the case that parameters are chosen so that

𝐼2=𝑘0,ln𝛼11/4+1/16+1/(2𝛼)1.(4.41)

Theorem 4.6. Assume 𝛼>1 and 𝑘>𝛼(1/(1/4+1/16+1/(2𝛼))1)1. Interval 𝐼2 given by (4.41) is not empty. For any 𝜏𝐼2, (𝐻2) holds and hence 𝜔1(𝜏)>0.

Proof. Assume 𝛼>1. Letting 1𝐺(𝛼)=4+1+11612𝛼𝛼,(4.42) then d1d𝛼𝐺(𝛼)=4𝛼2+11/16+1/(2𝛼)𝛼2=1𝛼211+8/𝛼+1>0.(4.43)𝐺(𝛼) is an increasing function of 𝛼 and 𝐺(1)=0. 𝐺(𝛼)>𝐺(1) implies that 1/4+1/16+1/(2𝛼)1/𝛼>0. Therefore 11>4+1+116>12𝛼𝛼.(4.44) This gives 1𝛼1>1/4+1/16+1/(2𝛼)1>0.(4.45) By assumption 𝑘>𝛼(1/(1/4+1/16+1/(2𝛼))1)1, we obtain 𝑘(𝛼1)𝛼>𝑘𝛼11/4+1/16+1/(2𝛼)1>1.(4.46) Noting that 𝐷=Δ=1 and recalling the definition of 𝜏𝑐 given in (3.2), 𝜏𝑐=ln𝑘(𝛼1)𝛼𝑘>ln𝛼11/4+1/16+1/(2𝛼)1>0.(4.47) Therefore, 𝐼2 given by (4.41) is not empty. For any 𝜏𝐼2, noting 𝑠+(𝜏)=1/(1+(𝛼Δ/𝑘)𝑒Δ𝜏) and Δ=1, we have 𝑠+(𝜏)[1/4+1/16+1/(2𝛼),1/(1+𝛼)/𝑘)[1/4+1/16+1/(2𝛼),1).
In what follows, we intend to show for any 𝜏𝐼2, or equivalently 𝑠+(𝜏)[1/4+1/16+1/(2𝛼),1), (𝐻2) holds. For any 𝑠+(𝜏)[1/4+1/16+1/(2𝛼),1), by (4.44), it follows that 𝑠+(𝜏)>1/𝛼 and so 𝛼𝑠+(𝜏)>1. Hence,
𝛽2(𝜏)𝑐2(𝜏)=2𝛼𝑠2+𝑠(𝜏)+1(𝜏)4211162𝛼𝛼𝑠+(𝜏)10.(4.48) For any 𝑠+(𝜏)[1/4+1/16+1/(2𝛼),1), we have 𝑠+(𝜏)>1/(4𝛼)+1/(16𝛼2)+1/(2𝛼), since 1/4>1/(4𝛼) and 1/16>1/(16𝛼2) imply that 14+1+116>12𝛼+4𝛼116𝛼2+1.2𝛼(4.49) Therefore, 𝑞2𝑝2(𝜏)+2𝛽(𝜏)=1𝑠+(𝜏)2𝛼𝑠2+𝑠(𝜏)+1(𝜏)4𝛼2116𝛼212𝛼>0.(4.50)
Condition (𝐻2) holds. By Lemma 4.4, 𝜔1(𝜏)>0.

Next, to determine whether (4.2) has a pair of pure imaginary eigenvalues, we consider

𝑐2(𝜏)+𝑞2𝜔2=𝛼𝑠+1(𝜏)𝑠+(𝜏)2+(1)2𝜔2=𝛼𝑠+1(𝜏)𝑠+(𝜏)2+𝜔2,𝜔𝑐(𝜏)(𝑝(𝜏)𝜔)+𝑞𝜔2𝜔𝛽(𝜏)=𝜔𝑐(𝜏)𝑝(𝜏)+𝑞21𝛽(𝜏)=𝜔𝛼𝑠2+(𝜏)𝜔2+𝛼1𝑠+,𝜔(𝜏)𝑐(𝜏)2=𝛽(𝜏)+𝑞𝜔(𝑝(𝜏)𝜔)𝛼𝑠+1(𝜏)𝑠+𝜔(𝜏)2𝛼1𝑠+(𝜏)+(1)𝜔𝜔𝑠+𝑠(𝜏)=𝛼+(𝜏)𝜔21𝑠+(𝜏)𝛼𝑠+1(𝜏)𝑠+(.𝜏)(4.51) We obtain

𝜔sin(𝜔𝜏)=𝜔2𝛼1/𝑠2+(𝜏)𝛼1𝑠+(𝜏)𝛼𝑠+(𝜏)1/𝑠+(𝜏)2+𝜔2,cos(𝜔𝜏)=𝛼𝑠+𝜔(𝜏)21𝑠+(𝜏)𝛼1/𝑠2+(𝜏)𝛼𝑠+(𝜏)1/𝑠+(𝜏)2+𝜔2.(4.52)

If there exists (𝜏,𝜔) satisfying (4.52), then (4.2) has a pair of pure imaginary roots ±𝑖𝜔. A necessary condition for (4.52) to have solutions is 𝛼𝑠2+(𝜏)1. Otherwise, 𝛼𝑠2+(𝜏)=1, and the second equation of (4.52) becomes cos(𝜔𝜏)=𝛼𝑠+(𝜏). However, for any 𝜏(0,𝜏𝑐), we have 𝛼𝑠+(𝜏)>1, since

𝛼𝑠+𝛼(𝜏)=1+𝛼𝑒𝜏>𝛼/𝑘=𝛼1+(𝛼/𝑘)𝑘(11/𝛼)=𝛼1+𝛼(11/𝛼)𝛼=1.(4.53) Hence, the second equation of (4.52) has no solution. Assume 𝛼𝑠2+(𝜏)1 for 𝜔0 and 𝜏[0,𝜏𝑐] and denote the right-hand sides of (4.52) by

1𝜔(𝜔,𝜏)=𝑐(𝜏)(𝑝(𝜏)𝜔)+𝑞𝜔2𝛽(𝜏)𝑐2(𝜏)+𝑞2𝜔2𝜔=𝜔2𝛼1/𝑠2+(𝜏)𝛼1𝑠+(𝜏)𝛼𝑠+(𝜏)1/𝑠+(𝜏)2+𝜔2,2𝜔(𝜔,𝜏)=𝑐(𝜏)2𝛽(𝜏)+𝑞𝜔(𝑝(𝜏)𝜔)𝑐2(𝜏)+𝑞2𝜔2=𝛼𝑠+𝜔(𝜏)21𝑠+(𝜏)𝛼1/𝑠2+(𝜏)𝛼𝑠+(𝜏)1/𝑠+(𝜏)2+𝜔2.(4.54) Define functions

𝜃1(𝜏)=arccos2𝜔1(𝜏),𝜏if2𝜔1[],𝜃(𝜏),𝜏1,12(𝜏)=arccos2𝜔2(𝜏),𝜏if2𝜔2([].𝜏),𝜏1,1(4.55)

Lemma 4.7. Assume 𝛼>(7+35)/2 and 𝑘>𝛼/(4𝛼1). For any 𝜏𝐼1 given by (4.19), there exists 𝜖𝑗>0 and 𝜃𝑗(𝜏) with 𝜖𝑗𝜃𝑗(𝜏)𝜋 (j=1,2) such that 𝜃sin𝑗(𝜏)+2𝑛𝜋=1𝜔𝑗,𝜃(𝜏),𝜏cos𝑗(𝜏)+2𝑛𝜋=2𝜔𝑗.(𝜏),𝜏𝑛=0,1,2,,(4.56)

Proof. For any 𝜏𝐼1, by Theorem 4.5, 𝜔1(𝜏)>0 and 𝜔2(𝜏)0. It is easy to see that 1(0,𝜏)=0 and lim𝜔+1(𝜔,𝜏)=. There are two roots of 1(𝑧,𝜏)=0, 𝑧1=0 and 𝑧2(𝜏)=1𝛼𝑠2+(𝜏)+𝛼1𝑠+.(𝜏)(4.57) Hence, 1(𝜔,𝜏)>0 for 0<𝜔<𝑧2(𝜏). For any 𝜏𝐼1, as shown in Theorem 4.5, 𝑠+(𝜏)[1/4𝛼,1/4+1/16+1/(2𝛼)]. This implies 𝑠+(𝜏)1/4𝛼>1/𝛼>1/𝛼. Therefore, 𝛼𝑠2+(𝜏)>1 and 𝛼𝑠+(𝜏)>1. The function 2(𝜔,𝜏) is monotonically increasing for 𝜔0, since 𝜕2(𝜔,𝜏)𝜕𝜔=𝛼𝑠+(𝜏)2𝜔𝛼𝑠+(𝜏)1/𝑠+(𝜏)2+𝜔2𝜔2𝜔21𝑠+(𝜏)𝛼1/𝑠2+(𝜏)𝛼𝑠+(𝜏)1/𝑠+(𝜏)2+𝜔22=2𝛼𝑠+(𝜏)𝜔𝛼𝑠+(𝜏)1/𝑠+(𝜏)2+1𝑠+(𝜏)𝛼1/𝑠2+(𝜏)𝛼𝑠+(𝜏)1/𝑠+(𝜏)2+𝜔22=2𝛼𝑠+(𝜏)𝜔𝛼1/𝑠2+𝑠(𝜏)2+(𝜏)𝛼1/𝑠2+(𝜏)+1𝑠+(𝜏)𝛼𝑠+(𝜏)1/𝑠+(𝜏)2+𝜔22=2𝛼𝑠+(𝜏)𝜔𝛼1/𝑠2+𝑠(𝜏)+(𝜏)𝛼𝑠+(𝜏)1𝛼𝑠+(𝜏)1/𝑠+(𝜏)2+𝜔22=2𝛼𝜔𝛼𝑠2+(𝜏)1𝛼𝑠+(𝜏)1𝛼𝑠+(𝜏)1/𝑠+(𝜏)2+𝜔220.(4.58) Since 𝑠+(𝜏)<1, 2(0,𝜏)=𝛼𝑠+(𝜏)1𝑠+(𝜏)𝛼1/𝑠2+(𝜏)𝛼𝑠+(𝜏)1/𝑠+(𝜏)2=𝛼𝑠+(𝜏)1𝑠+(𝜏)𝛼𝑠2+(𝜏)1<0.(4.59) Also, lim𝜔2(𝜔,𝜏)=𝛼𝑠+(𝜏)>1. Therefore, there exists a unique 𝜔=𝑙max(𝜏)=𝛼1/𝑠2+(𝜏)>0, such that 2(𝑙max(𝜏),𝜏)=1. Solving 2(𝑙max,𝜏)=1 for 𝑙max and noting that 𝛼1/𝑠2+(𝜏)0, it is easy to see that 𝑙max(𝜏)=𝛼1/𝑠2+(𝜏)𝑙2max𝛼𝑠+(𝜏)1=𝛼𝑠+(𝜏)1𝑠+1(𝜏)𝛼𝑠2++(𝜏)𝛼𝑠+1(𝜏)𝑠+(𝜏)2=𝛼𝑠+1(𝜏)𝛼𝑠2+(𝜏)𝛼𝑠+1(𝜏)+𝑠+(𝜏)+𝛼2𝑠2+1(𝜏)2𝛼+𝑠2+(𝜏)=𝛼2𝑠+(𝛼𝜏)𝑠+(𝜏)𝛼2𝑠2+(𝜏)+𝛼+𝛼2𝑠2+(1𝜏)2𝛼+𝑠2+(𝜏)=𝛼2𝑠+𝛼(𝜏)𝑠+1(𝜏)𝛼+𝑠2+=1(𝜏)𝛼𝑠2+(𝜏)𝛼𝑠+.(𝜏)1(4.60) Then, 2(𝜔,𝜏)1 for any 𝜔[0,𝑙max(𝜏)]. Since 𝑠+(𝜏)<1, 𝑙max(𝜏)<𝑧2(𝜏). Therefore, 1(𝜔,𝜏)>0 for any 𝜔[0,𝑙max(𝜏)]. Since 𝜔1(𝜏) is a positive root of 21(𝜔,𝜏)+22(𝜔,𝜏)=1, we have 2(𝜔1(𝜏),𝜏)1, which implies that 0<𝜔1(𝜏)𝑙max(𝜏)<𝑧2(𝜏). Therefore, 1(𝜔1(𝜏),𝜏)>0, and so 1(𝜔1(𝜏),𝜏)=12(𝜔1(𝜏),𝜏). In fact, 𝜔1(𝜏)<𝑙max(𝜏),(4.61) since 22𝑙max(𝜏),𝜏+21𝑙max(𝜏),𝜏=1+21𝑙max(𝜏),𝜏>1.(4.62) Thus, 𝜃1(𝜏) is defined and 0𝜃1(𝜏)𝜋. Since cos(𝜃1(𝜏)+2𝑛𝜋)=2(𝜔1(𝜏),𝜏), 𝜃sin1=(𝜏)+2𝑛𝜋1cos2𝜃1=(𝜏)+2𝑛𝜋122𝜔1(𝜏),𝜏=1𝜔1.(𝜏),𝜏(4.63) Hence, 𝜃1(𝜏) satisfies (4.56). From (4.61), 2(𝜔1(𝜏),𝜏)<2(𝑙max(𝜏),𝜏)=1, and so 𝜃1(𝜏)>0. Since 𝜃1(𝜏) is continuous on the interval 𝐼1 and 𝐼1 is closed, there exists 𝜖1>0 such that 𝜃1(𝜏)𝜖1. Similarly we can prove the existence of 𝜃2(𝜏).

Lemma 4.8. Assume 𝛼>1 and 𝑘>𝛼(1/(1/4+1/16+1/2𝛼)1)1. For any 𝜏𝐼2 given by (4.41), there exists 𝜖>0 and 𝜃1(𝜏) such that 𝜖𝜃1(𝜏)<𝜋 and 𝜃1(𝜏) satisfies (4.56) for 𝑗=1.

Proof. For any 𝜏𝐼2, by Theorem 4.6, only 𝜔1(𝜏)>0.
As in Lemma 4.7, we have 1(𝜔,𝜏)>0 for 0<𝜔<𝑧2(𝜏). For any 𝜏𝐼2, as shown in Theorem 4.6, 𝑠+(𝜏)[1/4+1/16+1/(2𝛼),1). Letting
1𝐺(𝛼)=4+1+11612𝛼𝛼,𝛼>1,(4.64) we have d1d𝛼𝐺(𝛼)=4𝛼2+11/16+1/(2𝛼)2𝛼𝛼=12𝛼1𝛼2+1/4+2𝛼𝛼>0.(4.65)𝐺(𝛼) is an increasing function and 𝐺(1)=0. Since 𝐺(𝛼)>𝐺(1), it follows that 1/4+1/16+1/(2𝛼)>1/𝛼. Therefore, for any 𝑠+(𝜏)>1/4+1/16+1/(2𝛼), we obtain 𝑠+(𝜏)>1/𝛼, or equivalently 𝛼>1/𝑠2+(𝜏). Since 𝜕2(𝜔,𝜏)𝜕𝜔=2𝛼𝑠+(𝜏)𝜔𝛼𝑠+(𝜏)1/𝑠+(𝜏)2+1𝑠+(𝜏)𝛼1/𝑠2+(𝜏)𝛼𝑠+(𝜏)1/𝑠+(𝜏)2+𝜔220,(4.66)2(𝜔,𝜏) is monotonically increasing for any 𝜔0. For any 𝑠+(𝜏)>1/4+1/16+1/(2𝛼), 𝑠+1(𝜏)4211162𝛼=𝑠2+𝑠(𝜏)+(𝜏)21=12𝛼2𝛼2𝛼𝑠2+(𝜏)𝛼𝑠+=1(𝜏)12𝛼𝛼𝑠2+(𝜏)1𝛼𝑠+(𝜏)𝛼𝑠2+=(𝜏)𝛼𝑠2+(𝜏)12𝛼1𝛼𝑠+(𝜏)1𝑠+(𝜏)𝛼𝑠2+(𝜏)1>0,(4.67) which implies that (𝛼𝑠+(𝜏)(1𝑠+(𝜏)))/(𝛼𝑠2+(𝜏)1)<1. Hence, 0>2(0,𝜏)=𝛼𝑠+(𝜏)1𝑠+(𝜏)𝛼1/𝑠2+(𝜏)𝛼𝑠+(𝜏)1/𝑠+(𝜏)2=𝛼𝑠+(𝜏)1𝑠+(𝜏)𝛼𝑠2+(𝜏)1>1.(4.68) For 𝜏[0,𝜏𝑐), lim𝜔2(𝜔,𝜏)=𝛼𝑠+(𝜏)>1, since 𝛼𝑠+𝛼(𝜏)=1+𝛼𝑒𝜏>𝛼/𝑘=𝛼1+(𝛼/𝑘)𝑘(11/𝛼)=𝛼1+𝛼(11/𝛼)𝛼=1.(4.69) As in the proof of Lemma 4.7, there exists a unique 𝑙max(𝜏)=𝛼1/𝑠2+(𝜏)>0 such that 2(𝑙max(𝜏),𝜏)=1. Then 𝑙max(𝜏)<𝑧2(𝜏). Therefore, 1(𝜔,𝜏)>0 for any 𝜔[0,𝑙max(𝜏)]. The rest of the proof is similar to the proof of Lemma 4.7. Furthermore, 𝜃1(𝜏)<𝜋, since 2(𝜔1(𝜏),𝜏)>1 for any 𝜔1(𝜏)[0,𝑙max(𝜏)).

Theorem 4.9. Consider system (2.3) with 𝐷=Δ=1. (1)Suppose 𝛼>(7+35)/2, 𝑘>𝛼/(4𝛼1), and 𝜏𝐼1 given by (4.22). For 𝜏𝐼1 and 𝑗=1,2, 𝜔𝑗(𝜏) is nonnegative and there exists 𝜖𝑗>0 and 𝜃𝑗(𝜏) such that 𝜖𝑗𝜃𝑗(𝜏)𝜋 and 𝜃𝑗(𝜏) satisfies (4.56). If there exists 𝑛0 such that 𝜃𝑗(𝜏)+2𝑛𝜋 intersects 𝜏𝜔𝑗(𝜏) at some 𝜏𝑗𝑛𝐼1, then (4.4) has a pair of pure imaginary eigenvalues 𝜆=±𝑖𝜔𝑗(𝜏𝑗𝑛). System (2.3) undergoes a Hopf bifurcation at 𝜏𝑗𝑛 provided 𝑑𝑅𝑒(𝜆(𝜏))/𝑑𝜏|𝜏=𝜏𝑗𝑛0.(2)Suppose 𝛼>1, 𝑘>𝛼(1/(1/4+1/16+1/(2𝛼))1)1, and 𝜏𝐼2 given by (4.41). For 𝜏𝐼2, only 𝜔1(𝜏) is positive and there exists 𝜖>0 and 𝜃1(𝜏) such that 𝜖𝜃1(𝜏)<𝜋 and 𝜃1(𝜏) satisfies (4.56) for 𝑗=1. If there exists 𝑛0 such that 𝜃1(𝜏)+2𝑛𝜋 intersects 𝜏𝜔1(𝜏) at some 𝜏1𝑛𝐼2, then (4.4) has a pair of pure imaginary eigenvalues 𝜆=±𝑖𝜔1(𝜏1𝑛). System (2.3) undergoes a Hopf bifurcation at 𝜏1𝑛 provided 𝑑𝑅𝑒(𝜆(𝜏))/𝑑𝜏|𝜏=𝜏1𝑛0.

Proof. Assume 𝐷=Δ=1 in system (2.3).Case 1. Suppose 𝜏𝐼1. By Theorem 4.5, 𝜔𝑗(𝜏)0 for 𝑗=1,2. By Lemma 4.7, there exists 𝜖𝑗>0 and 𝜃𝑗(𝜏) such that 𝜖𝑗𝜃𝑗(𝜏)𝜋 and 𝜃𝑗(𝜏) satisfies (4.56). Assume that there exists a positive integer 𝜏𝑗𝑛𝐼1 such that 𝜃𝑗(𝜏𝑗𝑛)+2𝑛𝜋=𝜏𝑗𝑛𝜔𝑗(𝜏𝑗𝑛) for some integer 𝑛0. Then system (4.52) has one solution (𝜏𝑗𝑛,𝜔𝑗(𝜏𝑗𝑛)). Equation (4.4) has a pair of pure imaginary eigenvalues 𝜆=±𝑖𝜔𝑗(𝜏𝑗𝑛).
In what follows, we show that the conditions required for a Hopf Bifurcation (see Theorem A.1 in the appendix) are satisfied by the linearization (3.3) of (2.3) at 𝐸+. In (A.1), choosing 𝜏 as the bifurcation parameter and letting
𝐷𝜏,𝑧𝑡=𝑧1000100011𝑧(𝑡)2𝑧(𝑡)3,𝐿(𝑡)𝜏,𝑧𝑡=1𝛼𝑥+𝛼𝑠+0𝛼𝑥+𝐷+𝛼𝑠+𝑘𝑦+𝑘𝑥+𝑧00Δ1𝑧(𝑡)2𝑧(𝑡)3+(𝑡)0000000𝑘𝑒Δ𝜏𝑦+𝑘𝑒Δ𝜏𝑥+𝑧1𝑧(𝑡𝜏)2𝑧(𝑡𝜏)3,(𝑡𝜏)(4.70) the linearization (3.3) of (2.3) at 𝐸+ is of the form (A.1). Taking 𝑎 to be any positive real number and 𝑏=1/2, hypothesis (𝑆1) in the Hopf Bifurcation Theorem holds, since |||||det𝑘=0𝐴𝑘(𝛼)𝑒𝜆𝑟𝑘(𝛼)|||||=||||||||||||||1det100010001=12,|||||det𝑘=0𝐴𝑘(𝛼)𝑒𝜆𝑟𝑘(𝛼)+01𝐴(𝛼,𝜃)𝑒𝜆𝜃|||||=||||||||||||||1d𝜃det100010001=12(4.71) for all 𝜏 and |Re𝜆|<𝑎.
The characteristic equation (4.4) of (3.3) at 𝐸+ has a pair of pure imaginary eigenvalues 𝜆=±𝜔1(𝜏𝑗𝑛) and no other root of (4.4) is an integral multiple of ±𝜔1(𝜏𝑗𝑛). Hence the hypothesis (𝑆2) in the Hopf Bifurcation Theorem holds. Therefore, (2.3) undergoes a Hopf bifurcation at 𝐸+ when 𝜏=𝜏𝑗𝑛 provided dRe(𝜆(𝜏))/d𝜏|𝜏=𝜏𝑗𝑛0.
Case 2. Suppose 𝜏𝐼2. By Theorem 4.5, only 𝜔1(𝜏)>0. By Lemma 4.8, there exists 𝜖>0 and 𝜃1(𝜏) such that 𝜖𝜃1(𝜏)<𝜋 and 𝜃1(𝜏) satisfies (4.56). Assume there exists 𝜏1𝑛𝐼2 such that 𝜃1(𝜏1𝑛)+2𝑛𝜋=𝜏1𝑛𝜔1(𝜏1𝑛) for some integer 𝑛0. Then system (4.52) has one solution (𝜏1𝑛,𝜔1(𝜏1𝑛)). Equation (4.4) has a pair of pure imaginary eigenvalues 𝜆=±𝑖𝜔1(𝜏1𝑛). The rest of the proof is similar to that of Case 1 when 𝑗=1.

Corollary 4.10. Consider system (2.3) with 𝐷=Δ=1. (1)Suppose 𝛼>(7+35)/2, 𝑘>𝛼/(4𝛼1), and 𝜏𝐼1 given by (4.22). For 𝜏𝐼1, 𝜔𝑗(𝜏) is nonnegative and there exists 𝜖𝑗>0 and 𝜃𝑗(𝜏) such that 𝜖𝑗𝜃𝑗(𝜏)𝜋 and 𝜃𝑗(𝜏) satisfies (4.56) for 𝑗=1,2. If there exists a positive integer 𝑛𝑗0 such that min𝜏𝐼1𝜏𝜔𝑗(𝜏)2𝑛𝑗𝜋 and max𝜏𝐼1𝜏𝜔𝑗(𝜏)>(2𝑛𝑗+1)𝜋, then 𝜃𝑗(𝜏)+2𝑛𝑗𝜋 intersects 𝜏𝜔𝑗(𝜏) at least once at some 𝜏𝑗𝑛𝑗𝐼1. System (2.3) undergoes a Hopf bifurcation at 𝜏𝑗𝑛𝑗 provided 𝑑𝑅𝑒(𝜆(𝜏))/𝑑𝜏|𝜏=𝜏𝑗𝑛𝑗0.(2)Suppose 𝛼>1, 𝑘>𝛼(1/(1/4+1/16+1/(2𝛼))1)1, and 𝜏𝐼2 defined in (4.41). For 𝜏𝐼2, only 𝜔1(𝜏) is positive. There exists 𝜖>0 and 𝜃1(𝜏) such that 𝜖𝜃1(𝜏)<𝜋 and 𝜃1(𝜏) satisfies (4.56) for 𝑗=1. If there exists a positive integer 𝑁0 such that max𝜏𝐼2𝜏𝜔1(𝜏)>(2𝑁+1)𝜋, then for any 0𝑛𝑁, 𝜃1(𝜏)+2𝑛𝜋 intersects 𝜏𝜔1(𝜏) at least once at some 𝜏1𝑛𝐼2. System (2.3) undergoes a Hopf bifurcation at 𝜏1𝑛 provided 𝑑𝑅𝑒(𝜆(𝜏))/𝑑𝜏|𝜏=𝜏1𝑛0.

Proof. Assume 𝐷=Δ=1 in system (2.3).Case 1. Suppose 𝜏𝐼1. By Theorem 4.5, 𝜔𝑗(𝜏)0 for 𝑗=1,2. By Lemma 4.7, there exists 𝜖𝑗>0 and 𝜃𝑗(𝜏) such that 𝜖𝑗𝜃𝑗(𝜏)𝜋 and 𝜃𝑗(𝜏) satisfies (4.56). Assume that there exists a positive integer 𝑛𝑗0 such that min𝜏𝐼1𝜏𝜔𝑗(𝜏)2𝑛𝑗𝜋 and max𝜏𝐼1𝜏𝜔𝑗(𝜏)>(2𝑛𝑗+1)𝜋. For such 𝑛𝑗, min𝜏𝐼1𝜏𝜔𝑗(𝜏)<𝜖𝑗+2𝑛𝑗𝜋𝜃𝑗(𝜏)+2𝑛𝑗𝜋2𝑛𝑗+1𝜋<max𝜏𝐼1𝜏𝜔𝑗(𝜏).(4.72) By the Mean Value Theorem, there exists 𝜏𝑗𝑛𝑗𝐼1 such that 𝜃𝑗(𝜏𝑗𝑛𝑗)+2𝑛𝑗𝜋=𝜏𝑗𝑛𝑗𝜔𝑗(𝜏𝑗𝑛𝑗). By Theorem 4.9. Case 1, the conclusion follows.Case 2. Suppose 𝜏𝐼2. By Theorem 4.5, only 𝜔1(𝜏)>0. By Lemma 4.8, there exists 𝜖>0 and 𝜃1(𝜏) such that 𝜖𝜃1(𝜏)<𝜋 and 𝜃1(𝜏) satisfies (4.56). Assume that there exists a positive integer 𝑁0 such that max𝜏𝐼2𝜏𝜔1(𝜏)>(2𝑁+1)𝜋. By (4.41), 0𝐼2. Therefore min𝜏𝐼2𝜏𝜔1(𝜏)=0. For 0𝑛𝑁, min𝜏𝐼2𝜏𝜔1(𝜏)<𝜖+2𝑛𝜋𝜃𝑖(𝜏)+2𝑛𝜋(2𝑛+1)𝜋<max𝜏𝐼2𝜏𝜔1(𝜏).(4.73) By the Mean Value Theorem, there exists 𝜏1𝑛𝐼2 such that 𝜃1(𝜏1𝑛)+2𝑛𝜋=𝜏1𝑛𝜔1(𝜏1𝑛). By Theorem 4.9. Case 2, the conclusion follows.

Corollary 4.11. Consider system (2.3) with 𝐷=Δ=1. Assume 𝛼>(7+35)/2 and 𝑘>𝛼/(4𝛼1). If 1/4+1/16+1/(2𝛼)>𝑘/(𝛼+𝑘), then 𝐼1=[0,ln((𝑘/𝛼)(4𝛼1))], where 𝐼1 was defined in (4.22). For any 𝜏𝐼1, 𝜔𝑗(𝜏) is nonnegative and there exists 𝜖𝑗>0 and 𝜃𝑗(𝜏) such that 𝜖𝑗𝜃𝑗(𝜏)𝜋 and 𝜃𝑗(𝜏) satisfies (4.56) for 𝑗=1,2. If there exists a positive integer 𝑁𝑗0 (𝑗=1,2) such that max𝜏𝐼1𝜏𝜔𝑗(𝜏)>(2𝑁𝑗+1)𝜋, then for any 0𝑛𝑁𝑗, 𝜃𝑗(𝜏)+2𝑛𝜋 intersects 𝜏𝜔𝑗(𝜏) at least once at some 𝜏𝑗𝑛𝐼1. System (2.3) undergoes a Hopf bifurcation at 𝜏𝑗𝑛 provided 𝑑𝑅𝑒(𝜆(𝜏))/𝑑𝜏|𝜏=𝜏𝑗𝑛0.

Proof. Assume 1/4+1/16+1/(2𝛼)>𝑘/(𝛼+𝑘). Then ln((𝑘/𝛼)(1/(1/4+1/16+1/(2𝛼))1))<0. By (4.22), 𝐼1=𝑘max0,ln𝛼11/4+𝑘1/16+1/(2𝛼)1,ln4𝛼1𝛼=𝑘0,ln𝛼4.𝛼1(4.74) For any 𝜏𝐼1, by Theorem 4.5, 𝜔𝑗(𝜏)0 for 𝑗=1,2. By Lemma 4.7, there exists 𝜖𝑗>0 and 𝜃𝑗(𝜏) such that 𝜖𝑗𝜃𝑗(𝜏)𝜋 and 𝜃𝑗(𝜏) satisfies (4.56). Noting 0𝐼1, min𝜏𝐼1𝜏𝜔𝑗(𝜏)=0. Assume there exists a positive integer 𝑁𝑗0 (𝑗=1,2) such that max𝜏𝐼1𝜏𝜔𝑗(𝜏)>(2𝑁𝑗+1)𝜋. For any 0𝑛𝑁𝑗, min𝜏𝐼1𝜏𝜔𝑗(𝜏)=02𝑛𝜋 and max𝜏𝐼1𝜏𝜔𝑗(𝜏)>(2𝑁𝑗+1)𝜋(2𝑛+1)𝜋. By Corollary 4.10, the conclusion follows.

5. Numerical Results

This section includes bifurcation diagrams involving the interior equilibrium 𝐸+ and numerical simulations of periodic solutions of the predator-prey model in the chemostat.

5.1. Variation of Eigenvalues

To study the stability switches of 𝐸+, DDEBIFTOOL (see [11, 12]) was chosen to illustrate how the real part of the eigenvalues of (4.2) changes as parameters 𝛼 and 𝜏 vary.

First fix parameters 𝐷=Δ=1, 𝑘=24, and 𝜏=0.5. Taking 𝛼 as the bifurcation parameter and varying it from 0 to 10, the real part of the eigenvalues with largest real part of (4.2) was plotted in Figure 1. At 𝛼1.15 and 𝛼1.5, there is either a zero eigenvalue or a pair of pure imaginary roots. For 𝛼(1.15,1.5), all eigenvalues have negative real parts. For example, taking 𝛼=1.3, Figure 2(a) shows that the eigenvalues of (4.2) with largest real parts (the ones in the circle) have negative real parts. Note that due to the scaling, the eigenvalues in the circle seem to be indistinguishable from zero. But in fact, they are a pair of complex eigenvalues with real parts slightly less than zero. DDEBIFTOOL can keep track of the occurrence of a pair of pure imaginary eigenvalues as 𝛼 varies in the neighborhood of 𝛼=1.5. Figure 2(b) clearly shows that there is a pair of pure imaginary eigenvalues. Hence, Hopf bifurcation is possible. Note that by continuation, the pair of eigenvalues with largest real parts in Figure 2(a) for 𝛼=1.3 becomes the pair of pure imaginary eigenvalues in Figure 2(b) for 𝛼1.5.

Finally fix all parameters as before and vary both 𝜏 and 𝛼. In Figure 3, we plot the Hopf bifurcation diagram in 𝛼 and 𝜏 parameter space. The curve at the left upper corner is 𝜏=𝜏𝑐. For any pair (𝛼,𝜏) below that curve, a coexistence equilibrium 𝐸+ exists (i.e., all components are positive). For any pair (𝛼,𝜏) on the closed curve, there is a Hopf bifurcation. Inside the closed curve, there is a periodic solution surrounding 𝐸+. For any (𝛼,𝜏) outside the closed curve and below 𝜏=𝜏𝑐, the coexistence equilibrium 𝐸+ is stable.

5.2. Simulations Demonstrating Hopf Bifurcations

In this section, we illustrate Theorem 4.9 for system (2.3). Take 𝐷=Δ=1 and let 𝜏 vary. We choose parameters 𝛼=100 and 𝑘=100 for Case 1 (see Figures 412), and 𝛼=2 and 𝑘=20 for Case 2 (see Figures 1318).

Case 1. Note that 𝐼1 is given by (4.22). Since ln((𝑘/𝛼)(1/(1/4+1/16+1/(2𝛼))1))=0.03, 𝑘max0,ln𝛼11/4+1/16+1/(2𝛼)1=0.(5.1) Also, ln(𝑘(4𝛼1)/𝛼)0.77. Therefore 𝐼1[0,0.77]. By Theorem 4.9, 𝜔𝑗(𝜏) is positive and 𝜃𝑗(𝜏) satisfies (4.56) for any 𝜏𝐼1 and 𝑗=1,2.
Figure 4 shows that 𝜃𝑗(𝜏) intersects 𝜏𝜔𝑗(𝜏) at some 𝜏𝑗0 with 𝜏100.022 and 𝜏200.48. We see that 𝜃𝑗(𝜏)+2𝑛𝜋 has no intersection with 𝜏𝜔𝑗(𝜏) for 𝑛2 and 𝑗=1,2. By Theorem 4.9, (4.4) has two distinct pairs of pure imaginary eigenvalues 𝜆=±𝑖𝜔𝑗(𝜏𝑗0). Next we need to check if Re(d𝜆(𝜏)/d𝜏)|𝜏=𝜏𝑗00.
As in Beretta and Kuang [13], we can define
𝑆𝑗𝑛𝜃(𝜏)=𝜏𝑗(𝜏)+2𝑛𝜋𝜔𝑗(𝜏),for𝑗=1,2,𝑛=0,1,2.(5.2) Any zero 𝜏𝑗𝑛 of 𝑆𝑗𝑛(𝜏) is an intersection of 𝜃𝑗(𝜏)+2𝑛𝜋 and 𝜏𝜔𝑗(𝜏) and vice versa. By (4.10) in Beretta and Kuang [13] and noting that (𝑞2𝑝2(𝜏)+2𝛼2(𝜏))24(𝛼2(𝜏)𝑐2(𝜏))>0, we have the relation signRed𝜆(𝜏)|||d𝜏𝜏=𝜏𝑗𝑛𝑞=±sign2𝑝2(𝜏)+2𝛼2(𝜏)2𝛼42(𝜏)𝑐2(𝜏)signd𝑆𝑗𝑛(𝜏)||||d𝜏𝜏=𝜏𝑗𝑛=±signd𝑆𝑗𝑛(𝜏)||||d𝜏𝜏=𝜏𝑗𝑛,(5.3) where we take + for 𝑗=1 and for 𝑗=2.
From Figure 5(a), it is observed that 𝑆1𝑛 has only one zero 𝜏100.022 at 𝑛=0 with sign{d𝑆10(𝜏)/d𝜏|𝜏=𝜏10}>0. Hence, sign{Re(d𝜆(𝜏)/d𝜏)|𝜏=𝜏10}>0. By Theorem 4.9, system (2.3) undergoes a Hopf bifurcation at 𝜏10. Similarly from Figure 5(b), 𝑆2𝑛 has only one zero 𝜏200.48 for 𝑛=0 and (2.3) undergoes a Hopf bifurcation at 𝜏20.
Next we used MATLAB to simulate solutions of model (2.3) for several values of 𝜏. For each fixed delay 𝜏, we chose initial data 𝑠(𝑡)=𝑠+(𝜏)0.01, 𝑥(𝑡)=𝑥+(𝜏)+0.01, and 𝑦(𝑡)=𝑦+(𝜏)+0.001 for 𝑡[𝜏,0]. From Figure 6, we can see that the equilibrium 𝐸+ is stable if 𝜏=0.02<𝜏10. As delay 𝜏 increases past 𝜏100.022, where a Hopf bifurcation occurs, a pair of complex eigenvalues of (4.4) enters the right-half plane. The equilibrium 𝐸+ loses its stability and a periodic solution bifurcates from 𝐸+ (see Figures 7 and 8). As we increase the delay further to 𝜏=0.4<𝜏20, the periodic solution still exists and remains stable (see Figures 9 and 10). However, as the delay 𝜏 increases further, past 𝜏20, the stable periodic solution disappears in a second Hopf bifurcation, and 𝐸+ regains stability (see Figure 11). We provide a bifurcation diagram illustrating the change in dynamics as 𝜏 varies (see Figure 12). For any 𝜏(𝜏10,𝜏20), there is an orbitally asymptotically stable periodic solution.

Case 2. Take 𝑘=20 and 𝛼=2. For such parameters, 𝐼2[0,0.85]. By Theorem 4.9, 𝜔1(𝜏) is positive and 𝜃1(𝜏) satisfies (4.56) for 𝑗=1 and 𝜏𝐼2. Figure 13 shows that 𝜃1(𝜏) intersects 𝜏𝜔1(𝜏) twice. To distinguish these intersections, denote them as 𝜏10,10.2 and 𝜏10,20.7. On the other hand, 𝜃1(𝜏)+2𝜋 has no intersection with 𝜏𝜔1(𝜏).
From Figure 14,
signd𝑆10(𝜏)||||d𝜏𝜏=𝜏10,1>0,signd𝑆11(𝜏)||||d𝜏𝜏=𝜏10,2<0.(5.4) By (5.3), signRed𝜆(𝜏)|||d𝜏𝜏=𝜏10,1>0,signRed𝜆(𝜏)|||d𝜏𝜏=𝜏10,2<0.(5.5) By Theorem 4.5, system (2.3) undergoes a Hopf bifurcation at 𝜏=𝜏10,1 and at 𝜏=𝜏10,2. For 𝜏 less than 𝜏10,1, the equilibrium 𝐸+ is asymptotically stable (see Figure 15). For 𝜏 greater than 𝜏10,1, but less than 𝜏10,2, there is an orbitally asymptotically stable periodic solution surrounding the equilibrium 𝐸+ (see Figures 16 and 17). At 𝜏=𝜏10,2, there is a second Hopf bifurcation, where the periodic solution coalesces with 𝐸+. For 𝜏>𝜏10,2, the periodic orbit no longer exists and 𝐸+ regains stability (see Figure 18) until it disappears when 𝜏>𝜏𝑐.

Appendix

Preliminary Results

To establish the existence of periodic solutions in autonomous delay differential equations, one of the simplest ways is through Hopf Bifurcation. Below is a general Hopf Bifurcation theorem for delay differential equations due to De Oliveira [14]. Before stating the theorem we require some notation.

Consider a one parameter family of neutral delay differential equations:

d𝐷d𝑡𝛼,𝑥𝑡𝑔𝛼,𝑥𝑡=𝐿𝛼,𝑥𝑡+𝑓𝛼,𝑥𝑡,𝛼,(A.1) where 𝐷,𝐿,𝑓, and 𝑔 are continuously differentiable in 𝛼 and 𝑥𝑡([𝑟,0],𝑛) (𝑟 is a constant), 𝑓(𝛼,0)=𝑔(𝛼,0), 𝜕𝑓(𝛼,0)/𝜕𝑥𝑡=𝜕𝑔(𝛼,0)/𝜕𝑥𝑡=0, 𝐷(𝛼,𝑥𝑡) and 𝐿(𝛼,𝑥𝑡) are linear in 𝑥𝑡, and

𝐷𝛼,𝑥𝑡=𝑘=0𝐴𝑘(𝛼)𝑥𝑡𝑟𝑘(+𝛼)01𝐿𝐴(𝛼,𝜃)𝑥(𝑡+𝜃)d𝜃,𝛼,𝑥t=𝑘=0𝐴𝑘(𝛼)𝑥𝑡𝑟𝑘+(𝛼)01𝐴(𝛼,𝜃)𝑥(𝑡+𝜃)d𝜃,(A.2) for 𝑥𝑡([𝑟,0],𝑛). Assume 𝛼, where 𝑟0(𝛼)=0, 𝑟𝑘(𝛼)(0,1], and 𝐴𝑘(𝛼), 𝐵𝑘(𝛼), 𝐴(𝛼,𝜃), and 𝐵(𝛼,𝜃) satisfy

𝑘=0||𝐴𝑘(||+||𝐵𝛼)𝑘(||+𝛼)01||||+||||𝐴(𝛼,𝜃)𝐵(𝛼,𝜃)d𝜃<.(A.3) It is easy to see that the characteristic matrix

Δ(𝛼,𝜆)=𝜆𝐷𝛼,𝑒𝜆𝐼𝐿𝛼,𝑒𝜆𝐼(A.4) is continuously differentiable in 𝛼 and Δ(𝛼,𝜆) is an entire function of 𝜆. Making the following assumptions on (A.1).

(𝑆1) There exist constants 𝑎>0, 𝑏>0 such that, for all complex values 𝜆 such that |𝑅𝑒𝜆|<𝑎 and all 𝛼, the following inequalities hold: |||||det𝑘=0𝐴𝑘(𝛼)𝑒𝜆𝑟𝑘(𝛼)||||||||||𝑏,det𝑘=0𝐴𝑘(𝛼)𝑒𝜆𝑟𝑘(𝛼)+01𝐴(𝛼,𝜃)𝑒𝜆𝜃|||||d𝜃𝑏.(A.5)(𝑆2) The characteristic equation detΔ(𝛼,𝜆)=0 has, for 𝛼=𝛼0, a simple purely imaginary root 𝜆0=𝑖𝑣0, 𝑣0>0, and no root of detΔ(𝛼0,𝜆)=0, other than ±𝑖𝑣0, is an integral multiple of 𝜆0.(𝑆3)𝑅𝑒(𝜕𝜆(𝛼0)/𝜕𝛼)0.

Now we are ready to state the Hopf bifurcation theorem for (A.1).

Theorem A.1 (Hopf bifurcation theorem, see Kuang [10, page 60]). In (A.1), assume that (𝑆1)-(𝑆3) hold. Then there is an 𝜖>0 such that, for 𝑎, |𝑎|𝜖, there are functions 𝛼(𝑎), 𝜔(𝑎), 𝛼(0)=𝛼0, 𝜔(0)=2𝜋/𝑣0, such that (A.1) has an 𝜔(𝛼)-periodic solution 𝑥(𝑎)(𝑡), that is continuously differentiable in 𝑡, and 𝑎 with 𝑥(0)=0. Furthermore, for |𝛼𝛼𝑜|<𝜖, |𝜔(2𝜋/𝑣0)|<𝜖, every 𝜔-periodic solution 𝑥(𝑡) of (A.1) with |𝑥(𝑡)|<𝜖 must be of this type, except for a translation in phase; that is, there exists 𝑎(𝜖,𝜖) and 𝑏 such that 𝑥(𝑡)=𝑥(𝑎)(𝑡+𝑏) for all 𝑡.

The following lemma is usually called the Fluctuation Lemma. For a proof, see Hirsh et al. [15].

Lemma A.2. Let 𝑓+ be a differentiable function. If liminf𝑡𝑓(𝑡)<limsup𝑡𝑓(𝑡), then there are sequences 𝑡𝑚 and 𝑠𝑚 such that for all 𝑚̇𝑓𝑡𝑚𝑡=0,𝑓𝑚limsup𝑡𝑓̇𝑓𝑠(𝑡)as𝑚,𝑚𝑠=0,𝑓𝑚liminf𝑡𝑓(𝑡)as𝑚,.(A.6)

The proof of the following useful lemma can be found in [16].

Theorem A.3. Let 𝑎(,) and 𝑓[𝑎,) be a differentiable function. If lim𝑡𝑓(𝑡) exists (finite) and the derivative function ̇𝑓(𝑡) is uniformly continuous on (𝑎,), then lim𝑡̇𝑓(𝑡)=0.

Acknowledgment

This research is partially supported by NSERC.