Hydrothermal liquefaction (HTL) is an effective method that can convert biomass into bio-crude, but direct use of bio-crude derived from biomass HTL remains a challenge due to the lower quality. In this study, bifunctional Ni/HZSM-5 catalysts and zinc hydrolysis were combined to produce upgraded bio-crude in an in-situ HTL process. The K2CO3 and HZSM-5 catalysts with different Ni loading ratios were tested. The effects of different catalysts on the yield and quality of bio-crude and gas were investigated. The results indicated that the catalysts improved bio-crude and gas yields, compared to pine sawdust liquefaction without catalyst. The catalysts reduced the contents of undesirable oxygenated compounds such as acids, ketones, phenols, alcohols and esters in bio-crude products while increased desirable hydrocarbons content. K2CO3 produced highest bio-crude yield and lowest solid residue yield among all catalysts. Compared to parent HZSM-5 catalyst, bifunctional Ni/HZSM-5 catalysts exhibited higher catalyst activity to improve quality of upgraded bio-crude due to its integration of cracking and hydrodeoxygenation reactions. 6%Ni/HZSM-5 catalyst produced the bio-crude with the highest hydrocarbons content at 11.02%. This catalyst can be a candidate for bio-crude production from biomass HTL.
1.
Introduction
There are some relationships between species such as predation, competition, mutualism and parasitism, while predation is a popular representative in ecosystem. Since Lotka [1] and Volterra [2] almost simultaneously proposed the Lotka-Volterra model of predation between two populations, many scholars have studied the dynamics of different types of predator-prey models (see [3,4,5,6]). They usually consider the response of the predator to the prey density changing through a functional response in the modelling of predator-prey model. A variety of models are generated due to the selection of different types of functional response functions. Holling type Ⅰ, Ⅱ and Ⅲ [7,8,9] were first proposed by Holling based on empirical field data from 1959 to 1966. The traditional predator-prey models assumed that the functional response depends only on the prey populations [10,11,12]. However, several biologists questioned functional response is determined solely by prey density. They suggested that predators might interfere with each other's foraging because predators share or compete for food with each other when predatos forage. Therefore, the functional response should depend not only on the prey density but also on the predator, see for instance [13,14,15,16,17,18,19,20,21,22]. The authors considered that the growth of population is on slow time scale (days or months) and established the ratio-dependent model. By virtue of Holling type Ⅱ function, Kesh et al.[19] proposed a ratio-dependent model and obtained the sufficient condition for permanent co-existence. Xu and Chen [20] and Tang and Chen [21] studied the local and global stability of two classes of ratio-dependent system with delay, respectively. Fan and Li [22] proposed a delayed discrete ratio-dependent model and obtained the sufficient conditions for the permanence.
The study on providing additional food for the predator is an important topic in ecological modeling, which leads to more complex dynamics for the predator-prey model and has received extensive attentions. In recent years, many scholars have focused on the work of providing additional food for the predator to protect species or control pest, for details to see references [23,24,25,26]. Assuming that the quality of the additional food remains constant, Srinivasu and Prasad [23] found that the growth of the predator could be enhanced by the provision of extra food for the predator. Basheer et al. [24] showed that providing extra food for the predator is effective in eliminating pests. Samaddar [25] and Wu [26] investigated the Hopf bifurcation of the model which took into account the provision of extra food for the predator and time delay.
Motivated by the view of Arditi and Ginzburg [27] and Srinivasu et al. [28], a ratio-dependent model which involves the assumption that the additional food is provided for the predator was first proposed by Kumar and Chakrabarty [29]. This extra food is considered to be available uniformly within the ecological domain [28]. The model is as follows
Where A is the additional food for the predator, the number of preys and predators at time T is respectively represented by N(T) and P(T), here N(0)>0,P(0)>0. The intrinsic growth rate of the prey is expressed by r and m′ stands for natural mortality of the predator, the carrying capacity of the prey is described by K. e1 and e2 denote the coefficients of predator capture on prey and additional food, n1 and n2 represent separately the nutritional value of the prey and additional food, h1 and h2 are handling time per predator per unit of the ratio NP prey biomass and per predator per unit of the ratio AP additional food biomass, respectively.
Another important aspect, the real ecosystem could not be denied the fact that the processes of biological development are not normally instantaneous on account of the interaction with environment or other species. Therefore, several biologists and mathematicians began to explore and study the time-delay models inspired by the dynamics of population models with time delay. From the biological and mathematical point of view, predator-prey models with delay [30,31,32,33,34,35] tend to exhibit more complex dynamics than those without that, such as local and global stability, Hopf bifurcation, limit cycle and so on. For example, the properties of the periodic solutions of the model with delay and impulse were obtained in [31]. Jiang [34] investigated the global asymptotical and the existence of Hopf bifurcation of a diffusive model with ratio-dependent function and delay. By studying the delay differential system, Tang and Zhang [35] showed that the stability of the model becomes unstable state when the time reaches a certain critical point, thus Hopf bifurcation occurs. Considering the maturation time of the prey, model (1.1) turns to the following version
It is widely known that human survival and development depend on natural resources. Biological resources have the most unique development mechanism due to renewability. Over-exploitation of biological resources will not only lead to a reduction in their quantity and quality, but also may drive species to extinction. Thus, taking the path of sustainable development becomes necessary to maintain ecological balance and meets the material needs of the people. From an economic perspective, the impact of harvesting effort on ecosystem guarantees that the net economic revenue is equal to the difference between total revenue and total cost which may influence on the harvesting activities. There has been a growing literature for the modeling and analysis of bioeconomic systems. The existence of steady state was discussed in a bioeconomic model with ratio-dependent [36]. The dynamics of the delay and diffusion terms of a bioeconomic plankton model was investigated by Zhao et al. [37]. In particular, an increasing number of bioeconomic models are expressed by differential-algebraic equations [38,39,40,41,42,43,44]. A prey-predator economic model with time-delay and stage structure was proposed by Zhang et al. [38]. Through detailed proof, the results show that the model undergoes three different bifurcational phenomena. Chen [41] established a biological economic model which suggested that the model undergoes flip bifurcation and Neimark-Sacker bifurcation.
To extend the model above, a bioeconomic model with time delay incorporating two differential equations and one algebraic equation is written as
From an economic perspective, the impact of harvesting effort on ecosystem is expressed as: net economic revenue (NER) = total revenue (TR) - total cost (TC). TR =p1q1E(t)P(t) and TC =wE(t), where p1 is the price per unit predator biomass, q1 and E(t) stand for the harvesting capacity coefficient and harvesting effort of the predator, respectively. w represents the cost per unit predator harvesting efforts. Q means the NER. We only consider the harvesting of the predator, the catch rate function h(t) is written as h(t)=q1E(t)P(t) which meets the catch-per-unit-effort hypothesis [45]. All the parameters in model (1.3) are positive. The initial conditions for model (1.3) are given N[−τ,0]∈C+([−τ,0],R+). If P(0) is provided, E(0) can be expressed as Q/(p1q1P(0)−w) thus it is content with P(0)>w/p1q1.
Significantly, a real-life application is motivated by data from Katz [46]. We consider that prey N is represented by barnacles Balanus balanoides and predator P is expressed as snails Urosalpinx cinerea. The data strongly supports the ratio-dependent form N/P much better than that only depends on prey N. Incorporating additional food A (Ostrea gigas thunberg) into our model, our target is to address our understanding on the consequences of providing additional food for the predator on the dynamics, which brings out an explicit link between practical biological control and theoretical fruits. Besides, we investigate the optimal harvesting strategy of snails Urosalpinx cinerea due to its use as a kind of Chinese medical herbs.
The rest of this manuscript is emerged as follows. Section 2 is devoted to discussing the local steady-state and the Hopf bifurcation by investigating the characteristic equation. In Section 3, the main results including the properties of the bifurcating periodic solutions and the formulas of the direction of Hopf bifurcation are investigated. The optimal harvesting problem of model (1.3) without time delay is analyzed in Section 4. Four numerical examples are displayed in Section 5 to explain our findings. The brief and powerful biological discussion and conclusions are provided in Section 6. The paper end with Appendices A, B and C to show some proof and calculations.
2.
Local stability of equilibria and Hopf bifurcation
Model (1.3) is simplified and analyzed by using the following rescaling
After nondimensionalizing the model (1.3) for convenience, we obtain
where c=e1r, α=n1h2n2h1, ξ=γAK, γ=n2e2n1e1, b=n1rh1, m=m′r, q=q1r, p=p1Ke1h1r.
Through calculation, the internal equilibrium of the model (2.1) could be written as P∗0(x∗,y∗,E∗), where
and x∗ is the positive solution of Ax4+Bx3+Cx2+Dx+F=0 where
Remark 2.1. By Descartes sign rule [47], the possible positive roots of equation Ax4+Bx3+Cx2+Dx+F=0 are as follows.
(ⅰ) If any one of the conditions (a), (b), (c) and (d) holds, the equation exists one positive root
(ⅱ) If any one of the conditions (e), (f), (g), (h), (i) and (j) holds, the equation exists two positive roots or no roots
(ⅲ) If any one of the conditions (k), (l), (m) and (n) holds, the equation exists three positive roots or one positive root
(ⅳ) If the condition (p) holds, the equation exists four positive roots or two positive roots or no roots
According to the following theorem, we know the conditions of the Hopf bifurcation occurs in model (2.1).
Theorem 2.1. If the assumptions A1+A4>0,A3>0,A2+A3>0 and A2−A3<0 are satisfied, then when τ<τ0, equilibrium P∗0(x∗,y∗,E∗) of model (2.1) is stable; equilibrium P∗0(x∗,y∗,E∗) of Eq (2.1) is unstable when τ>τ0; when τ=τ0, model (2.1) undergoes a Hopf bifurcation.
The detailed proving process of the Theorem is given in Appendix A.
3.
Stability and direction of the Hopf bifurcation
We investigate the direction of Hopf bifurcation and the properties of the bifurcating periodic solutions at the positive equilibrium in this section.
Theorem 3.1. If μ2>0 (μ2<0), the model undergoes a supercritical (subcritical) Hopf bifurcation at P∗(x∗1,y∗1,E∗1); moreover, if β2<0 (β2>0) then the bifurcating periodic solutions are stable (unstable); if T2>0 (T2<0), the periodic of bifurcating periodic solutions increase (decrease). Here
The detailed proof of the Theorem is given in Appendix B.
4.
Optimal harvesting
In order to achieve species persistence and profit maximization, optimal harvesting problem for model (1.3) without time delay is discussed in this section. We consider instantaneous annual discount rate δ, we give a continuous time stream of revenues J as follows
The control constrain on this interval 0<E(t)<Emax. Here, Emax stands for the maximum harvesting effort. The optimal prey and predator densities and the corresponding optimal harvesting effort are represented by Noptimal, Poptimal and Eoptimal, respectively. We are committed to calculating the optimal harvesting effort Eoptimal which satisfies the following formula
where U presents the control set written as
The calculation procedure of Noptimal, Poptimal and Eoptimal are shown in Appendix C.
5.
Numerical simulation
In this section, we will employ several specific examples to simulate the solutions to model (1.3), and verify the analytical results of the existence of Hopf bifurcation. Our analytical results illustrate that Hopf bifurcation arises in model (1.3) when the value of bifurcation parameter τ crosses critical values τ=τk, where
and ω10 is the root of equation ω14+(A21−2A2−A24)ω12+(A22−A23)=0. Here, bifurcation parameter τ is the maturation time of prey. In order to illustrate the Hopf bifurcation of (1.3). We first choose the parameter values as follows: r=1.01, K=0.91, e2=1.01, h1=0.99, h2=0.01, n1=0.85, n2=0.98, A=0.98, m′=1.99, q1=0.91, p1=360, w=1.01, Q=0.51. The scientific significance of these parameters are shown in Table 1, under the above parameters, model (1.3) becomes the following model (5.1)
Through maple software, we get P1(N∗,P∗,E∗)=(0.7191,0.2939,0.0054) and P∗1(x∗1,y∗1,E∗1)= (0.7903,0.7248,0.0108) are respectively the equilibrium of models (5.1) and (A.2) with above parameters.
Substituting the value of the positive equilibrium P∗1(x∗1,y∗1,E∗1) in the expression of A1,A2,A3 and A4 to calculate the values as follows: A1=0.81, A2=−0.16, A3=0.73 and A4=0.79 which surely satisfy the conditions of Theorem 2.1 (that is A1+A4=1.60>0, A2+A3=0.56>0 and A2−A3=−0.89<0). In addition, we further substitute A1,A2,A3 and A4 to equation (2.8) to calculate ω10=0.74, then substituting ω10 in the expression of τk to yield τ0=1.86. A series of calculations show that g20=−12.70+3.11i, g11=−4.43−10.67i, g02=11.05−3.71i, g21=−106.05+74.57i, c1(0)=−97.31−43.88i, Re{λ′(τk)}=1.01, Im{λ′(τk)}=2.35, μ2=96.35>0, β2=−194.62<0, T2=134.80>0. Hence, by Theorem 3.1, the periodic increase, and the direction is supercritical, and the periodic solutions are stable. By Matlab we will reveal how the model can arise Hopf bifurcation by choosing different time delays τ in the following figures.
When τ=1.63<τ0=1.86, model (5.1) satisfies the condition of Theorem 2.1. So P1(N∗,P∗,E∗)=(0.7191,0.2939,0.0054) is asymptotically stable. Figure 1 reveals the time evolution and the phase trajectory of model (5.1) for τ=1.63. It is obvious that model (5.1) converges to the positive equilibrium.
When τ=1.87>τ0=1.86, model (5.1) satisfies the condition of Theorem 2.1. So P1(N∗,P∗,E∗)=(0.7191,0.2939,0.0054) is unstable, i.e., a family of periodic solutions are bifurcated from P1(N∗,P∗,E∗). The periodic solutions and its stability are shown in Figure 2.
Combined with Figures 1 and 2, we can see that the size of maturation time τ affects the dynamic behaviors of the model, see Figure 3. This means that if time delay τ exceeds critical value, bifurcation behavior occurs at the positive equilibrium P1(N∗,P∗,E∗) of model (1.3).
We choose the parameter values as follows: r=0.2, K=0.1, e2=0.7, h1=0.95, h2=0.8, n1=0.1, n2=3, A=0.01, m′=0.9, q1=5, p1=1, w=0.01, δ=0.01. (Noptimal,Poptimal) and Eoptimal as capture rate e1 increase are shown by (a) in Figure 4. It shows that with the increase of capture rate e1, the optimal prey population decreases initially and increases afterwards, optimal predator population gradually decreasing, and the optimal harvesting effort gradually decreases. Besides, we choose the parameter values as follows: r=1, K=0.45, e1=60, e2=0.2, h1=1, h2=0.02, n1=1.5, n2=1, m′=0.01, q1=1.44, p1=0.02, w=1.01, δ=0.1. (Noptimal,Poptimal) and Eoptimal as additional food A increase are shown by (b) in Figure 4. It displays that with the increase of additional food A, the corresponding optimal prey and predator populations gradually increase, and the optimal harvesting effort gradually decreases.
6.
Discussion and conclusions
In our paper, based on a predator-prey model with additional food proposed by Kumar et al. [29], we incorporate time delay and harvesting into the above model and establish a differential-algebraic bio-economic model. Through a series of analytical analysis we obtain the results about the existence of positive equilibrium, the conditions of steady-state and Hopf bifurcation, and the direction of bifurcation as well as the stability of bifurcating periodic solutions. Firstly, the local stability condition of positive equilibrium of model (2.1) with time delay is A2−A3>0, while that without time delay are A1+A4>0 and A2+A3>0. Theorem 2.1 implies that the presence of the mature delay can destabilize model (2.1). When τ<τ0, equilibrium P∗0(x∗,y∗,E∗) of model (2.1) is stable. The equilibrium is unstable when τ>τ0; when τ=τ0, model (2.1) has a Hopf bifurcation near the equilibrium. Besides, Theorem 3.1 displays that the direction of the bifurcation is determined by μ2, the stability of bifurcating periodic solutions is determined by β2, and the period of the bifurcating periodic solutions is determined by T2. Last, optimal harvesting strategies for model (1.3) without time delay are investigated. We achieve optimal harvesting equilibrium (Noptimal,Poptimal) and optimal harvesting effort Eoptimal.
The numerical simulation of the model is carried out through Maple and Matlab software, and we find that the results of the numerical simulation agree with the theoretical results in Theorems 2.1 and 3.1. The influence of the bifurcation parameter τ on the dynamics of prey, predator and harvesting effort is discussed. The experimental results illustrate that the positive equilibrium of the model is locally asymptotically stable when bifurcation parameter τ<1.86, see Figures 1. Conversely, equilibrium P1(N∗,P∗,E∗) is unstable and there exist bifurcating periodic solutions which are stable {when bifurcation parameter τ>1.86, see Figure 2. The bifurcation diagram of (1.3) when τ=1.86 is presented in Figure 3, and the direction of bifurcation is supercritical. This also illustrates the influence of time delay τ which is one of the causes of population size fluctuation on the Hopf bifurcation of model (1.3), } i.e. the presence of the mature delay can destabilize model (1.3). Furthermore, we display optimal harvesting strategies in Figure 4, and depict the change of optimal harvesting equilibrium and optimal harvesting effort with respect to the rate of predator capture on prey e1 and extra food A, respectively. The phenomenon in Figure 4 (a) is expected due to the stronger impact of the coefficient of predator capture on prey e1. From Figure 4 (a), we know that optimal harvesting effort approaches zero when the coefficient of predator capture on prey e1=0.21. This means that when the coefficient of predator capture on prey e1 is beyond the level 0.21, harvesting predators should not be considered in order to prevent extinction. Figure 4 (b) shows that the optimal harvesting effort gradually decreases, while optimal prey and predator populations gradually increase, since additional food A distracts the predator from attacking the prey. At the same time, in order to prevent harmful prey from growing to uncontrollable numbers, the harvesting effort on predator should be reduced.
There are many topics that require us to refine and further analyze our model. On the one hand, the discrete time delay is considered in this study of Hopf bifurcation while more rich dynamic behaviors are generated by incorporating the distributed delays. On the other hand, the influence of random factors could not be ignored in which we introduce white noise into our model which involving the multiplicative random excitation and additive random excitation to analyze the stochastic P-bifurcation and stochastic D-bifurcation. The model is displayed as follows
where ξ(t),η(t) represent the multiplicative random excitation and additive random excitation respectively. The above mentions make our model more interesting and complicated and remain to be done in the future.
Use of AI tools declaration
The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.
Acknowledgments
The authors thank the editor and referees for their careful reading and valuable comments. The work is supported by the National Natural Science Foundation of China (No.12101211) and the Program for Innovative Research Team of the Higher Education Institution of Hubei Province (No.T201812) and the Scientific Research Project of Education Department of Hubei Province (No.B2020090) and the Program of Graduate Education Innovation of Hubei Minzu University (No.MYK2023042).
Conflict of interest
The authors declare that there is no conflict of interest.
Appendix A
In order to linearize model (2.1), we consider the following linear transformation
Thus model (2.1) could be rewritten as
The internal equilibrium of the model (A.2) becomes P∗1(x∗1,y∗1,E∗1), we transform the equilibrium into zero by linear transformation x1=x∗1+x2, y1=y∗1+y2, E1=E∗1+I(x2,y2), where
Model (A.2) is expressed as the following equation according to the above transformations
By calculation, the Jacobian matrix of the model (A.3) at (0,0) is represented as follows
The formula for the determination of the characteristic equation gives that λ2+A1λ+A2+(A3+A4λ)e−λτ=0, where
When τ=0, the characteristic equation is as follows
By Routh-Hurwitz criterion, we know that P∗1(x∗1,y∗1,E∗1) is locally asymptotically stable provided that the conditions A1+A4>0 and A2+A3>0.
When τ>0, the characteristic equation is yielded to
Suppose that iω1 is a solution of (A.5), one gets
Separating the real and imaginary parts of Eq (A.6), we obtain
Furthermore,
Let u=ω21. Then, Eq (A.8) becomes
If A1>0,A3>0,A2+A3>0 and A2−A3<0, then Eq (A.8) exists a unique positive real root ω10, hence (A.5) has a pair of pure imaginary roots ±iω10 at the critical value τk>0, which could be denoted as
If A2−A3>0, (A.5) has no real roots and P∗1(x∗1,y∗1,E∗1) is asymptotical stability for any τ>0.
We have to prove that ddτReλ(τk)>0 is true for guaranteeing the existence of Hopf bifurcation. Consider the following equation
Using a computation process similar to that of Tang [48] and Cooke [49], one yields
Since A2+A3>0 and A2−A3<0, we know that f(u) has a unique positive real root u0=ω210. In addition it can be shown that f′(u0)>0 (that is f′(ω210)>0). Combined with the above analysis, we have the following theorem.
Appendix B
Let u1(t)=x2(τt), u2(t)=y2(τt), τ=τk+μ (μ∈R), then, μ=0 is the Hopf bifurcation value. The functional differential equation is considered as follows
where u(t)=(u1(t),u2(t))∈R2 and Lμ:C→R2,F:R×C→R2. Let us define
and
where
in which μ1=(x∗1+y∗1+αξ), ν1=(pqy∗1−w) and ϕ(θ)=(ϕ1(θ),ϕ2(θ))∈C.
There exists a 2×2 matrix function η(θ,μ) (θ∈[−1,0]) such that
At this point, bounded variation function could be chosen as
Here, δ(θ) is the Dirac function.
For ϕ∈C1([−1,0],R2), we define
Then, Eq (B.1) is equivalent to the following equation
The bifurcating periodic solution u(t,μ) of Eq (B.1) is affected by parameter ε, solution u(t,μ(ε)) of Eq (B.1) exists amplitude o(ε), nonzero exponential β(ε) (β(0)=0) and period T(ε). Under the above assumptions, μ, T, and β have the following expansion form
Next, we compute the coefficients μ2, T2, β2. For ψ∈C1([0,1],(R2)∗), assign the conjugate operator A∗ of A is as follows
the adjoint bilinear form is given by
Obviously, ±iω10τk are both eigenvalues of A(0) and A∗. Now we calculate the eigenvector of A(0) corresponding to iω10τk. Let q(θ)=(1,Δ)Teiω10τkθ, then, A(0)q(θ)=iω10τkq(θ). Combining Eq (B.6) and the definition of A(0), it is easy to get
we further obtain
By a similar argument as that in the above part, the eigenvector of A∗ corresponding to −iω10τk could be achieved. Let q∗(s)=D(1,Δ∗)eiω10τks, we also have
From Eq (B.12) one gets
Hence,
Apparently, q∗(s) q(θ) meets ⟨q∗(s),q(θ)⟩=1 and ⟨q∗(s),ˉq(θ)⟩=0. When μ=0, let μt be the solution of Eq (B.9), we set
Therefore,
here, the local coordinates of center manifold C0 corresponding to q and q∗ are expressed as z and ˉz. Note that W is real if ut is real, thus only real solutions are considered. For solution ut∈C0, since μ=0,
that is
where
Comparing Eqs (B.18) and (B.20), one has
where q(θ)=(1,Δ)Teiω10τkθ, furthermore
Combining Eqs (B.3) and (B.4), we obtain
Comparing with the corresponding coefficients of the Eq (B.23), it yields to
In order to deduce g21, now we have to calculate W20(θ) and W10(θ). It follows from Eqs (B.9) and (B.18) that
where
Consider on the center manifold C0 close enough to the origin, we have
Substituting Eq (B.30) into (B.28) and comparing the coefficients on z2 and zˉz, it easy to see
For θ∈[−1,0), from Eq (B.28) we can get
Comparing the coefficients of Eq (B.29) gives that
We have from Eqs (B.31) and (B.33) that
Noticing that q(θ)=q(0)eiω10τkθ, then
where M1=(M(1)1,M(2)1)∈R2 is a constant vector. By the same method as that in the above part we can calculate
where M2=(M(1)2,M(2)2)∈R2 is also a constant vector.
Calculating constant vectors M1 and M2 is our main work in the following discussion. It follows from Eqs (B.28) and (B.29) that
where
and
According to the definition of A and Eq (B.31) we have
and
where η(θ)=η(0,θ). Substituting Eqs (B.35), (B.36) and (B.37) into (B.31), we notice that
it is easy to see
Hence, we further get
and
From Eqs (B.35) and (B.36), we can determine the values of W20(θ) and W11(θ). g21 be represented by the parameters of Eq (B.1). Therefore, we further calculate the values of c1(0), μ2, β2 and T2, as follows
Appendix C
Before finding the maximum value of J(E), the concrete Hamiltonian function is given
where λi=λi(t),i=1,2, are adjoint variables. The condition that the Hamiltonian function H satisfy is presented by
The adjoint variables λ1 and λ2 satisfy the following adjoint equations according to Pontryagin's Maximum Principle.
In order to find the solutions λ1(t), λ2(t) of Eq (C.6), we delete λ2 in Eq (C.6) to obtain a second order differential equation with respect to λ1
where
Then, we obtain λ1(t)=O1e−δt, where O1=C1δ2−I1δ+I2. Set up the equation in a similar way as above d2λ2dt2+I1dλ2dt+I2λ1=C2e−δt, we get λ2(t)=O2e−δt, where O2=C2δ2−I1δ+I2, C2=−Ep1q1(δ+rNK−e21h1NP(1+e1h1NP+e2h2AP)2). Substituting λ2(t) into (C.5) yields
Therefore, the values of (Noptimal,Poptimal) and Eoptimal can be obtained by Eq (C.9) and solving the biological equilibrium.
The optimal harvesting effort at any time is determined by the following optimal harvesting solution
where Emin is the minimum harvesting effort.