
Citation: Tara P. Hurst, Caroline Coleman-Vaughan, Indu Patwal, Justin V. McCarthy. Regulated intramembrane proteolysis, innate immunity and therapeutic targets in Alzheimer’s disease[J]. AIMS Molecular Science, 2016, 3(2): 138-157. doi: 10.3934/molsci.2016.2.138
[1] | Hongying Shu, Wanxiao Xu, Zenghui Hao . Global dynamics of an immunosuppressive infection model with stage structure. Mathematical Biosciences and Engineering, 2020, 17(3): 2082-2102. doi: 10.3934/mbe.2020111 |
[2] | Hongbin Guo, Michael Yi Li . Global dynamics of a staged progression model for infectious diseases. Mathematical Biosciences and Engineering, 2006, 3(3): 513-525. doi: 10.3934/mbe.2006.3.513 |
[3] | Yijun Lou, Bei Sun . Stage duration distributions and intraspecific competition: a review of continuous stage-structured models. Mathematical Biosciences and Engineering, 2022, 19(8): 7543-7569. doi: 10.3934/mbe.2022355 |
[4] | Jia Li . Malaria model with stage-structured mosquitoes. Mathematical Biosciences and Engineering, 2011, 8(3): 753-768. doi: 10.3934/mbe.2011.8.753 |
[5] | Wei Feng, Michael T. Cowen, Xin Lu . Coexistence and asymptotic stability in stage-structured predator-prey models. Mathematical Biosciences and Engineering, 2014, 11(4): 823-839. doi: 10.3934/mbe.2014.11.823 |
[6] | Sophia R.-J. Jang . Discrete host-parasitoid models with Allee effects and age structure in the host. Mathematical Biosciences and Engineering, 2010, 7(1): 67-81. doi: 10.3934/mbe.2010.7.67 |
[7] | Shunyi Li . Hopf bifurcation, stability switches and chaos in a prey-predator system with three stage structure and two time delays. Mathematical Biosciences and Engineering, 2019, 16(6): 6934-6961. doi: 10.3934/mbe.2019348 |
[8] | Zhisheng Shuai, P. van den Driessche . Impact of heterogeneity on the dynamics of an SEIR epidemic model. Mathematical Biosciences and Engineering, 2012, 9(2): 393-411. doi: 10.3934/mbe.2012.9.393 |
[9] | Asma Alshehri, John Ford, Rachel Leander . The impact of maturation time distributions on the structure and growth of cellular populations. Mathematical Biosciences and Engineering, 2020, 17(2): 1855-1888. doi: 10.3934/mbe.2020098 |
[10] | Jianxin Yang, Zhipeng Qiu, Xue-Zhi Li . Global stability of an age-structured cholera model. Mathematical Biosciences and Engineering, 2014, 11(3): 641-665. doi: 10.3934/mbe.2014.11.641 |
Jellyfish plays an important role in the marine ecosystems as a keystone species and a potential resource for human consumption [1]. The amount of jellyfish has been significantly increasing in many waters since 1980s [2]. Jellyfish can be found in many regions worldwide such as Japan [3], the China Seas [4], the Mediterranean Sea [5], Taiwan [6], Southampton Water and Horsea Lake, England [7]. It can survive in a wide range of water temperatures $ (0- 36\; ^{\circ}C) $ and salinities $ (3- 36\%) $ [8,9].
Jellyfish has a complex life history with several different phases: planula, polyp, strobila, ephyra and medusa [10]. The polyp and medusa are two main stages of the life cycle of jellyfish. Medusae are dioecious, the sperm combines with egg to form a planula, which normally settles to the bottom and then occur metamorphosis of planula into tentacles-ring polyp (or scyphistoma) [11]. For Aurelia aurita jellyfish, the scyphistoma produces external outgrowths asexually by budding, the vitally asexual reproduction of polyp ($ 94\% $), stolon ($ 5\% $) and podocysts ($ 1\% $) [3]. The scyphistoma changes into strobila (strobilating polyp) through strobilation, which is asexual reproduction by division into segments developing into ephyra. After liberating from strobila, the ephyra becomes adult medusa. In addition, strobila reverts into the initial scyphistoma [11]. Since jellyfish has a distinct mobility patterns in different phases of its life history, it is interesting to take these facts into account for model formulation.
Temperature has a great effect on variations of jellyfish populations [12] as the asexual reproduction rate and strobilation rate depend on the functions of temperatures [11]. Global warming has affected the increase of jellyfish populations because it might cause the distribution, growth, and ephyrae production of medusae [13]. The rapid strobilation might be proceeded at the warmer temperature, but the continuous high temperature results in the fewer budding and increased mortality [6]. Hence the population explosions of polyps and medusae might be caused at the appropriate increase of temperature, but rising temperatures lead to the decreased populations.
Many approaches for jellyfish have been developed to discuss the nature of the correlations between environmental indices and population abundance [6,14]. In particular, mathematical modeling is one of the important tools in analyzing the dynamical properties in aquatic systems. In [15], Oguz et al. presented food web model with an anchovy population and bioenergetics-based weight growth model governed by system of differential equations. In [16], Melica et al. conducted that the dynamics of polyps population by the logistic model. In [11], Xie et al. proposed the following two-dimensional dynamic model of scyphozoan jellyfish:
$ \begin{equation} \begin{aligned} \frac{dP}{dt} = & \alpha(T)P+s_1\gamma M-d_1P-d_2P-b_1P^2, \\ \frac{dM}{dt} = & s_2\beta(T)nP-d_3M-d_4M-b_2M^2, \end{aligned} \end{equation} $ | (1.1) |
where $ P(t) $ and $ M(t) $ are the population sizes of polyps and medusae at time $ t $, respectively. For system (1.1), by using the Bendixson-Dulac's negative criterion and Poincare-Bendixson Theorem, the conditions for the global asymptotical stability of the equilibria $ E_0 $, $ E_1 $ and $ E^* $ are given. The effects of temperature, substrate and predation on the population sizes of scyphozoan were investigated by numerical simulations.
Although multiple progresses have been seen in the above work of [11], for system (1.1), it is assumed that each population preserves an equal density dependent rate and each individual has the same opportunity to compete for their common resources during the whole life history. Unfortunately, this is not realistic due to the life history of jellyfish which has a diverse mobility body structures in different stages. The immature stages of jellyfish are much weaker than the mature stages and so they cannot compete for their common resources. Jellyfish reaches maturity after surviving the immature stages. Therefore, it is realistic and interesting for us to construct the stage-structured jellyfish model that exhibits a diversity between these different stages.
Recently, population dynamic models with stage structure and time delays have attracted more attention from authors [17,18,19,20,21,22,23,24]. For instance, Aiello and Freedman proposed a stage-structured model of single species containing of the immature and mature stages and using a discrete time delay taken from birth to maturity [18]. Liu et al. showed that the global stability for the two competitive Lotka-Volterra system with time delay that denotes the time taken from birth to maturity. They proposed that the stage structure is one of the main reasons that cause permanence and extinction for the two competitive system [23]. There have been an increasing interest and progress in the study of the above stage-structured models which assume all individuals are in the same species that require the analogous amount of time to become mature at the same age. Unlike birds and mammals, jellyfish species have the different mobility shapes in the distinctive stages of its life cycle. Thus, the previous methods and techniques cannot be applied exactly to our system because we classify the single species jellyfish into two-stage structure. Mathematically, the proposed model has two delay terms and the equations are matched with each other, which is not similar with the previous models [18,23,24].
In this paper, we will propose a time-delayed jellyfish model with stage structure and will investigate how the stage structure parameters and temperature affect the dynamical behaviors of system (2.2). Our main purpose is to study the population dynamics of jellyfish for the largest surviving probability as well as for final population numbers. To find the largest surviving probability, we will take the global asymptotical behaviors of the model by applying the monotone dynamical properties (for reference, see [25] and [26,p. 90]).
This paper is organized as follows: in Section 2, we propose the model and show that the solutions are positiveness and ultimately bounded. The main results of this paper are presented in Section 3. In Section 4, we perform numerical simulations to explore the effects of two delays and temperature on the dynamics. Section 5 is the brief discussion of our results.
The life history of jellyfish is divided into two main stages; polyp and medusa. The larval stage of polyp is planula and the elementary phase of medusa is ephyra. Let $ P(t) $ and $ M(t) $ be the population size or number of polyps and medusae at time $ t $, respectively. The model is based on the following assumptions and the diagram in Figure 1:
$ \rm(A1) $ $ \tau_1 $ is the length of the stage from the young polyp to the mature polyp. The immature polyp reproduces asexually at time $ t-\tau_1 $ and surviving from time $ t-\tau_1 $ to $ t $ is $ e^{-(d_1+d_2)\tau_1} $.
$ \rm(A2) $ $ \tau_2 $ is the time lag taken from the developed polyp to ephyra (incipient medusa), i.e., the developed polyp reaches ephyra after surviving this stage. Denote $ \tau = \max\lbrace{\tau_1, \tau_2}\rbrace $.
$\rm (A3) $ Its maturity is denoted by $ \tau = \max\lbrace{\tau_1, \tau_2}\rbrace, \tau > 0 $.
$ \rm(A4) $ Each population competes for their common resources.
$ \rm(A5) $ Each population has its own natural death rate, the mortality of polyp is varied by the factors of silt coverage or nudibranch consumption while that of medusa is because of different types of predators.
By the preceding assumptions, we get the following polyp-medusa population with stage structure:
$ \begin{equation} \begin{array}{rcl} \frac{dP}{dt}& = &\alpha(T)e^{-(d_1+d_2)\tau_1}P(t-\tau_1)+s_1\gamma M-d_1P-d_2P-b_1P^2, \\[6pt] \frac{dM}{dt}& = &s_2\beta(T)ne^{-(d_3+d_4)\tau_2}P(t-\tau_2)-d_3M-d_4M-b_2M^2. \end{array} \end{equation} $ | (2.1) |
As pointed out in [11], $ \alpha(T) $ denotes the asexual reproduction rate affected by temperature, involving budding, stolon and podocyst et al., $ s_1 $ is the survival and metamorphosis rate of planula, $ \gamma $ represents the sexual reproduction rate, $ b_1 $ and $ b_2 $ denote the density-dependent rates of polyps and medusae respectively, $ s_2 $ is the survival and development rate of ephyra, $ \beta(T) $ is the strobilation rate affected by temperature and $ n $ is strobilation times. Assuming that the death rate of the immature population is proportional to the existing immature population with proportionality constants $ d_i > 0 $, i = 1, 2, 3, 4. The loss of polyp is either due to natural death rate $ d_1 $ or due to the factors of silt coverage $ d_2 $ and $ \tau_1 $ is the time taken from the immature polyp to the mature one; thus $ e^{-(d_1+d_2)\tau_1} $ is the survival probability of each immature polyp to reach the mature one. The death of medusa is either from natural fatality rate $ d_3 $ or because of the predations $ d_4 $ and $ \tau_2 $ is the time length between the developed polyp and ephyra (incipient medusa); thus $ e^{-(d_3+d_4)\tau_2} $ is the survival rate of each developed polyp to reach the ephyra (incipient medusa) population. As it takes a few days in the larval stage of jellyfish life, the permanence and extinction criteria for the stage structured model are independent in this larval stage.
For the goal of simplicity, we denote $ a = \alpha(T), $ $ b = s_1\gamma, $ $ d = d_1+d_2 $, $ c = s_2\beta(T)n, $ $ d_* = d_3+d_4 $. Thus the following system can be obtained from system (2.1).
$ \begin{equation} \begin{array}{ll} \frac{dP}{dt} & = ae^{-\zeta_1}P(t-\tau_1)+bM-dP-b_1P^2, \\[6pt] \frac{dM}{dt} & = ce^{-\zeta_2}P(t-\tau_2)-d_*M-b_2M^2, \end{array} \end{equation} $ | (2.2) |
where $ \zeta_1 = d\tau_1 $ and $ \zeta_2 = d_*\tau_2 $. Denote $ \zeta_1 $ and $ \zeta_2 $ the degrees of the stage structure.
Let $ X = \mathcal{C}([-\tau, 0], \mathbb{R}^2) $ be the Banach space of all continuous function from $ [-\tau, 0] $ to $ \mathbb{R}^2 $ equipped with the supremum norm, where $ \tau = \max\lbrace{\tau_1, \tau_2}\rbrace $. By the standard theory of functional differential equations (see, for example, Hale and Verduyn Lunel [27]), for any $ \psi\in\mathcal{C}([-\tau, 0], \mathbb{R}^2) $, there exists a unique solution $ Y(t, \psi) = (P(t, \psi), M(t, \psi)) $ of system (2.2); which satisfies $ Y_0 = \psi $.
For system (2.2), we consider the initial conditions to either the positive cone $ X^+ = \{\psi\in X|\mbox{ $\psi_i(\theta)\ge 0$ for all $\theta\in[-\tau, 0]$, $i = 1, 2$ }\} $ or the subset of $ X^+ $ of functions which are strictly positive at zero, $ X_0^+ = \lbrace{\psi\in X^+|\psi_i(0) > 0, i = 1, 2}\rbrace $.
Lemma 2.1. For equation
$ \begin{equation} \frac{d\overline{W}}{dt} = ae^{-\zeta_1}\overline{W}(t-\tau_1)+\frac{bce^{-\zeta_2}}{d_*}\overline{W}(t-\tau_2)-\frac{B}{2}\overline{W}^2, \end{equation} $ | (2.3) |
where $ a, b, c, d_*, B > 0 $, $ \tau = \max\lbrace{\tau_1, \tau_2}\rbrace, \tau > 0 $ and $ \overline{W}(0) > 0 $ and $ \overline{W}(\theta)\ge 0 $, $ \theta\in[-\tau, 0] $, we have: $ \lim_{t\to\infty}\overline{W}(t) = \frac{2d_*ae^{-\zeta_1}+2bce^{-\zeta_2}}{d_*B} $ if $ d_*ae^{-\zeta_1}+bce^{-\zeta_2} > 0 $.
Proof. By using the similar argument of Lu et al. [28,Proposition 1], we will prove that $ \overline{W}(t) > 0 $, for all $ t\ge0 $. Otherwise, there exists some constant $ \acute{t}_0 > 0 $ such that $ \min \lbrace{\overline{W}(\acute{t}_0)}\rbrace = 0. $ Let $ t_0 = \inf \lbrace{\acute{t}_0: \overline{W}(\acute{t}_0) = 0}\rbrace. $ Then we have that $ \min \lbrace{\overline{W}(t_0)} = 0\rbrace $ and $ \min \lbrace{\overline{W}(t)}\rbrace > 0 $, $ \forall t \in [0, t_0) $. From system (2.3), we have
$ \begin{equation} \begin{aligned} \overline{W}(t)& = \overline{W}(0)e^{-\int_0^t \frac{B}{2}\overline{W}(\vartheta)d\vartheta}+ae^{-\zeta_1}\int_0^t \overline{W}(\omega-\tau_1)e^{-\int_\omega^t \frac{B}{2}\overline{W}(\vartheta)d\vartheta}d\omega\\ &+\frac{bce^{-\zeta_2}}{d_*}\int_0^t \overline{W}(\omega-\tau_2)e^{-\int_\omega^t \frac{B}{2}\overline{W}(\vartheta)d\vartheta}d\omega. \end{aligned} \end{equation} $ | (2.4) |
Incorporation initial conditions and Eq (2.4), we get $ \overline{W}(t_0) > 0 $, contradicting $ \min \lbrace{\overline{W}(t_0)}\rbrace = 0 $. Consequently, $ \overline{W}(t) > 0 $ for all $ t \ge 0 $.
Let $ \overline{W}^* = \frac{2d_*ae^{-\zeta_1}+2bce^{-\zeta_2}}{d_*B} $ denotes the unique positive equilibrium of system (2.3). Denote $ u(t) = \overline{W}(t)-\overline{W}^* $, thus system (2.3) takes the form as
$ \begin{equation} \begin{aligned} \frac{du}{dt} = ae^{-\zeta_1}u(t-\tau_1)+\frac{bce^{-\zeta_2}}{d_*}u(t-\tau_2)-\frac{B}{2}u^2(t)-B\overline{W}^*u(t). \end{aligned} \end{equation} $ | (2.5) |
Constructing the Lyapunov functional
$ \begin{equation*} \begin{aligned} V(u, u_\tau) = \frac{1}{2}u^2(t)+\frac{1}{2}ae^{-\zeta_1}\int_{t-\tau_1}^{t}u^2(s)ds+\frac{1}{2}\frac{bce^{-\zeta_2}}{d_*}\int_{t-\tau_2}^{t}u^2(s)ds, \end{aligned} \end{equation*} $ |
we have
$ \begin{equation*} \begin{aligned} \left.\frac{dV}{dt}\right|_{(2.5)} = &ae^{-\zeta_1}u(t)u(t-\tau_1)+\frac{bce^{-\zeta_2}}{d_*}u(t)u(t-\tau_2)-\frac{B}{2}u^3(t)-B\overline{W}^*u^2(t)+\frac{1}{2}ae^{-\zeta_1}u^2(t)\\ &-\frac{1}{2}ae^{-\zeta_1}u^2(t-\tau_1)+ \frac{1}{2}\frac{bce^{-\zeta_2}}{d_*}u^2(t)-\frac{1}{2}\frac{bce^{-\zeta_2}}{d_*}u^2(t-\tau_2)\\ \le&\frac{1}{2} ae^{-\zeta_1}u^2(t)+\frac{1}{2} ae^{-\zeta_1}u^2(t-\tau_1)+\frac{1}{2}\frac{bce^{-\zeta_2}}{d_*}u^2(t)+\frac{1}{2}\frac{bce^{-\zeta_2}}{d_*}u^2(t-\tau_2)-\frac{B}{2}u^3(t)\\ &-B\overline{W}^*u^2(t)+\frac{1}{2}ae^{-\zeta_1}u^2(t)-\frac{1}{2}ae^{-\zeta_1}u^2(t-\tau_1)+ \frac{1}{2}\frac{bce^{-\zeta_2}}{d_*}u^2(t)-\frac{1}{2}\frac{bce^{-\zeta_2}}{d_*}u^2(t-\tau_2)\\ = &ae^{-\zeta_1}u^2(t)+\frac{bce^{-\zeta_2}}{d_*}u^2(t)-B\overline{W}^*u^2(t)-\frac{B}{2}u^3(t)\\ = &(ae^{-\zeta_1}+\frac{bce^{-\zeta_2}}{d_*}-\frac{B}{2}\overline{W}^*)u^2(t)-(\overline{W}^*+u(t))\frac{B}{2}u^2(t)\\ = &-\frac{B}{2}\overline{W}(t)u^2(t)\le 0, \end{aligned} \end{equation*} $ |
which is negative definite and $ \left.\frac{dV}{dt}\right|_{(2.5)} = 0 $ if and only if $ u = 0 $. By Lyapunov-LaSalle invariance principle ([29,Theorem 2.5.3]), we get $ \lim_{t\to\infty}\overline{W}(t) = \overline{W}^* = \frac{2d_*ae^{-\zeta_1}+2bce^{-\zeta_2}}{d_*B} $, this proves Lemma 2.1.
Lemma 2.2. Given system (2.2), then:
(I) Under the initial conditions, all the solutions of system (2.2) are positive for all $ t\ge 0 $.
(II) Solutions of system (2.2) are ultimately bounded.
Proof. (I) We start with proving the positivity of solutions by using the similar argument of Lu et al. [28,Proposition 1]. We will prove that $ P(t) > 0 $, $ M(t) > 0 $ for $ t\ge 0 $. Otherwise, there exists some constant $ \tilde{t}_0 > 0 $ such that $ \min\lbrace{P(\tilde{t}_0), M(\tilde{t}_0)\rbrace = 0} $. Let $ t_0 = \inf\lbrace{\tilde{t}_0:P(\tilde{t}_0) = 0, M(\tilde{t}_0) = 0}\rbrace $. Then we have that $ \min\lbrace{P(t_0), M(t_0)}\rbrace = 0 $ and $ \min\lbrace{P(t), M(t)}\rbrace > 0 $, $ \forall t\in[0, t_0) $. From system (2.2), we have
$ \begin{equation} \left\{ \begin{aligned} P(t) = &P(0)e^{-\int_{0}^{t}(d+b_1P(\eta))d\eta}+ae^{-\zeta_1} \int_{0}^{t}P(\kappa-\tau_1)e^{-\int_{\kappa}^{t}(d+b_1P(\eta))d\eta} d\kappa\\ &+b\int_{0}^{t}M(\kappa)e^{-\int_{\kappa}^{t}(d+b_1P(\eta))d\eta} d\kappa, \\ M(t) = &M(0)e^{-\int_{0}^{t}(d_*+b_2M(\eta))d\eta} +ce^{-\zeta_2} \int_{0}^{t}P(\kappa-\tau_2)e^{-\int_{\kappa}^{t}(d_*+b_2M(\eta))d\eta} d\kappa. \end{aligned} \right. \end{equation} $ | (2.6) |
Incorporation initial conditions and Eq (2.6), we obtain $ P(t_0) > 0 $ and $ M(t_0) > 0 $, contradicting $ \min\lbrace{P(t_0), M(t_0)}\rbrace = 0 $. Consequently, $ P(t) > 0 $, $ M(t) > 0 $ for all $ t\ge 0 $.
(II) We show that the boundedness of solutions as follows.
Let $ W = \frac{d_*}{b}P+M $. By system (2.2), we have
$ \begin{equation*} \begin{aligned} \left.\frac{dW}{dt}\right|_{(2.2)} = &ae^{-\zeta_1}\frac{d_*}{b}P(t-\tau_1)+ce^{-\zeta_2}P(t-\tau_2)-\frac{dd_*}{b}P-\frac{d_*b_1}{b}P^2-b_2M^2\\ = &ae^{-\zeta_1}[\frac{d_*}{b}P(t-\tau_1)+M(t-\tau_1)]-ae^{-\zeta_1}M(t-\tau_1)+\frac{bce^{-\zeta_2}}{d_*}[\frac{d_*}{b}P(t-\tau_2)+M(t-\tau_2)]\\ &-\frac{bce^{-\zeta_2}}{d_*}M(t-\tau_2)-\frac{dd_*}{b}P-\frac{bb_1}{d_*}(\frac{d_*}{b}P)^2-b_2M^2\\ \le&ae^{-\zeta_1}W(t-\tau_1)+\frac{bce^{-\zeta_2}}{d_*}W(t-\tau_2)-B[(\frac{d_*}{b}P)^2+M^2], \end{aligned} \end{equation*} $ |
where $ B: = \min{\{\frac{bb_1}{d_*}, b_2}\} $.
$ \frac{dW}{dt}\le ae^{-\zeta_1}W(t-\tau_1)+\frac{bce^{-\zeta_2}}{d_*}W(t-\tau_2)-\frac{B}{2}W^2, $ |
where $ -2((\frac{d_*}{b}P)^2+M^2)\le -(\frac{d_*}{b}P+M)^2 $.
Consider the equation
$ \begin{equation} \frac{d\overline{W}}{dt} = ae^{-\zeta_1}\overline{W}(t-\tau_1)+\frac{bce^{-\zeta_2}}{d_*}\overline{W}(t-\tau_2)-\frac{B}{2}\overline{W}^2. \end{equation} $ | (2.7) |
By using Lemma 2.1 and Comparison Theorem, we get
$ \limsup_{t\to\infty}W(t)\le \frac{2d_*ae^{-\zeta_1}+2bce^{-\zeta_2}}{d_*B} $, which implies $ P(t) $ and $ M(t) $ are ultimately bounded. This completes the proof of Lemma 2.2.
The equilibria $ (P, M) $ of system (2.2) satisfies the following system
$ \begin{equation} \begin{aligned} ae^{-\zeta_1}P+bM-dP-b_1P^2 = 0, \\ ce^{-\zeta_2}P-d_*M-b_2M^2 = 0. \end{aligned} \end{equation} $ | (2.8) |
System (2.2) has the equilibria $ E_0 = (0, 0) $ for all parameter values and $ E_1 = (\frac{ae^{-\zeta_1}-d}{b_1}, 0) $ if $ ae^{-\zeta_1}-d > 0 $ and $ c = 0 $.
Since Eq (2.8) can be rewritten as
$ \begin{equation} \begin{aligned} (ae^{-\zeta_1}-d-b_1P)P& = -bM, \ \ ce^{-\zeta_2}P& = (d_*+b_2M)M. \end{aligned} \end{equation} $ | (2.9) |
When $ PM\neq 0 $, from Eq (2.9) it follows that
$ \begin{equation} \frac{ae^{-\zeta_1}-d-b_1P}{ce^{-\zeta_2}}+\frac{b}{d_*+b_2M} = 0. \end{equation} $ | (2.10) |
Further, substituting $ P = \frac{(d_*+b_2M)M}{ce^{-\zeta_2}} $ into Eq (2.10) and we get
$ \frac{ae^{-\zeta_1}-d-b_1 \frac{(d_*+b_2M)M}{ce^{-\zeta_2}}}{ce^{-\zeta_2}}+\frac{b}{d_*+b_2M} = 0. $ |
Set
$ F(M): = \frac{ae^{-\zeta_1}-d-b_1 \frac{(d_*+b_2M)M}{ce^{-\zeta_2}}}{ce^{-\zeta_2}}+\frac{b}{d_*+b_2M}, $ |
thus $ F(M) $ is a decreasing function of $ M $ for any $ M > 0 $.
Noting that the continuity and monotonicity of $ F(M) $ and that $ F(+\infty) < 0 $, furthermore since one can get $ F(0) > 0 $ provided that
$ \begin{equation} (ae^{-\zeta_1}-d)d_*+bce^{-\zeta_2} > 0, \ \ c\neq 0 \end{equation} $ | (2.11) |
hold true, therefore we conclude that system (2.2) admits a unique positive equilibrium given Eq (2.11) is satisfied.
The purpose of this section is to study the global stability of system (2.2).
Now we consider the local stability of the equilibria. The characteristic equation of system (2.2) takes the form as follows;
$ \begin{equation*} \det(\lambda I-G-H_1e^{-\lambda \tau_1}-H_2e^{-\lambda \tau_2}) = 0, \end{equation*} $ |
where
$ \begin{equation*} G = \begin{pmatrix} -d-2b_1P &b\\ 0&-d_*-2b_2M \end{pmatrix}, H_1 = \begin{pmatrix} ae^{-\zeta_1} &0\\ 0&0 \end{pmatrix}, H_2 = \begin{pmatrix} 0 &0\\ ce^{-\zeta_2}&0 \end{pmatrix}. \end{equation*} $ |
Lemma 3.1. Suppose that $ (ae^{-\zeta_1}-d)d_*+bce^{-\zeta_2} < 0 $, then the equilibrium $ E_0 = (0, 0) $ of system (2.2) is locally asymptotically stable.
Proof. The characteristic equation of system (2.2) at the equilibrium $ E_0 $ is as follows:
$ \begin{equation} C(\lambda): = (\lambda+d_*)(\lambda+d-ae^{-\zeta_1-\lambda\tau_1})-bce^{-\zeta_2-\lambda\tau_2} = 0. \end{equation} $ | (3.1) |
To show that it is asymptotically stable under assumption $ (ae^{-\zeta_1}-d)d_*+bce^{-\zeta_2} < 0 $, we just need to prove that the solutions of the characteristic equation $ C(\lambda) = 0 $ must have negative real parts. Let $ \lambda = u+iv $, where $ u $ and $ v $ are real numbers. Denote
$ \begin{equation*} \begin{aligned} A_1 = &u+d_*, \qquad &B_1 = &v, \\ A_2 = &u+d-ae^{-\zeta_1-u\tau_1}\cos(v\tau_1), \qquad &B_2 = &v+ae^{-\zeta_1-u\tau_1}\sin(v\tau_1), \\ C_1 = &bce^{-\zeta_2-u\tau_2}\cos(v\tau_2), \qquad &C_2 = &-bce^{-\zeta_2-u\tau_2}\sin(v\tau_2). \end{aligned} \end{equation*} $ |
Substituting $ \lambda $ by $ u+iv $ into Eq (3.1)
$ \begin{equation*} \begin{aligned} A_1A_2-B_1B_2 = C_1, \quad A_1B_2+A_2B_1 = C_2. \end{aligned} \end{equation*} $ |
Then
$ \begin{equation} (A_1A_2)^2+(B_1B_2)^2+(A_1B_2)^2+(A_2B_1)^2 = (C_1)^2+(C_2)^2. \end{equation} $ | (3.2) |
Assume that $ u\ge 0 $, then we get
$ A_1\ge d_* > 0, \quad A_2\ge d-ae^{-\zeta_1} > 0. $ |
Hence
$ \begin{equation} \begin{aligned} (A_1A_2)^2 > ((d-ae^{-\zeta_1})d_*)^2. \end{aligned} \end{equation} $ | (3.3) |
$ \begin{equation*} (A_1A_2)^2+(B_1B_2)^2+(A_1B_2)^2+(A_2B_1)^2 \ge (A_1A_2)^2 > ((d-ae^{-\zeta_1})d_*)^2. \end{equation*} $ |
From Eq (3.2), we obtain
$ \begin{equation*} \begin{aligned} (C_1)^2+(C_2)^2&\ge (A_1A_2)^2 > ((d-ae^{-\zeta_1})d_*)^2\\ (bce^{-\zeta_2})^2 (\cos^2(v\tau_2)+\sin^2(v\tau_2))&\ge (A_1A_2)^2 > ((d-ae^{-\zeta_1})d_*)^2\\ (bce^{-\zeta_2})^2&\ge (A_1A_2)^2 > ((d-ae^{-\zeta_1})d_*)^2. \end{aligned} \end{equation*} $ |
Hence Eq (3.3) contradicts to the assumption $ (ae^{-\zeta_1}-d)d_*+bce^{-\zeta_2} < 0 $, thus $ u < 0 $, which means $ \lambda $ must have negative real parts. This proves Lemma 3.1.
Lemma 3.2. Suppose that $ ae^{-\zeta_1}-d > 0 $ and $ c = 0 $, then the equilibrium $ E_1 = (\frac{ae^{-\zeta_1}-d}{b_1}, 0) $ of system (2.2) is locally asymptotically stable.
Proof. The characteristic equation of system (2.2) at the equilibrium $ E_1 $ is
$ \begin{equation} \begin{aligned} X(\lambda): = (\lambda+d_*)(\lambda-d+2ae^{-\zeta_1}-ae^{-\zeta_1-\lambda\tau_1}) = 0. \end{aligned} \end{equation} $ | (3.4) |
Then $ \lambda = -d_* $ is a negative root of the equation $ X(\lambda) = 0 $. Let $ \lambda-d+2ae^{-\zeta_1}-ae^{-\zeta_1-\lambda\tau_1} = 0 $; then if the root is $ \lambda = u+iv $, we have $ u+2ae^{-\zeta_1}-d-ae^{-\zeta_1-u\tau_1}\cos(v\tau_1) = 0 $. Assume that $ u\ge 0 $, then $ u+2ae^{-\zeta_1}-d-ae^{-\zeta_1-u\tau_1}\cos(v\tau_1)\ge ae^{-\zeta_1}-d > 0 $ is a contradiction, hence $ u < 0 $. This shows that all the roots of $ X(\lambda) = 0 $ must have negative real parts, and therefore $ E_1 $ is locally asymptotically stable. This proves Lemma 3.2.
Lemma 3.3. Suppose that $ (ae^{-\zeta_1}-d)d_*+bce^{-\zeta_2} > 0 $ and $ c\neq0 $, then the equilibrium $ E^* = (P^*, M^*) $ of system (2.2) is locally asymptotically stable.
Proof. The characteristic equation of system (2.2) at the equilibrium $ E^* $ is
$ \begin{equation*} \begin{aligned} (\lambda+d+2b_1P^*-ae^{-\zeta_1-\lambda\tau_1})(\lambda+d_*+2b_2M^*)-bce^{-\zeta_2-\lambda\tau_2} = 0. \end{aligned} \end{equation*} $ |
To show that it is asymptotically stable under $ (ae^{-\zeta_1}-d)d_*+bce^{-\zeta_2} > 0 $, we just need to prove that the solutions of the characteristic equation must have negative real parts. Let $ \lambda = u+iv $ where $ u $ and $ v $ are real numbers. Denote
$ \begin{aligned} D_1 = &u+d+2b_1P^*-ae^{-\zeta_1-u\tau_1}\cos(v\tau_1), &E_1 = &v+ae^{-\zeta_1-u\tau_1}\sin(v\tau_1), \\ D_2 = &u+d_*+2b_2M^*, &E_2 = &v, \\ F_1 = &bce^{-\zeta_2-u\tau_2}\cos(v\tau_2), &F_2 = &-bce^{-\zeta_2-u\tau_2}\sin(v\tau_2). \end{aligned} $ |
Substituting $ \lambda $ by $ u+iv $ into the above equation.
$ \begin{equation*} \begin{aligned} D_1D_2-E_1E_2 = F_1, \quad D_1E_2+D_2E_1 = F_2. \end{aligned} \end{equation*} $ |
Then
$ \begin{equation} \begin{aligned} (D_1D_2)^2+(E_1E_2)^2+(D_1E_2)^2+(D_2E_1)^2 = (F_1)^2+(F_2)^2. \end{aligned} \end{equation} $ | (3.5) |
Assume that $ u\ge 0 $, then we get
$ \begin{equation*} \begin{aligned} D_1&\ge d+2b_1P^*-ae^{-\zeta_1} > d+2b_1 \frac{ae^{-\zeta_1}-d}{b_1}-ae^{-\zeta_1} = ae^{-\zeta_1}-d > 0, \\ D_2&\ge d_*+2b_2M^* = d_*+2b_2\frac{b_1(P^*)^2-(ae^{-\zeta_1}-d)P^*}{b_1}\\ & > d_*+2b_2 \frac {b_1(\frac{ae^{-\zeta_1}-d}{b_1})^2-(ae^{-\zeta_1}-d)(\frac{ae^{-\zeta_1}-d}{b_1})}{b_1} = d_* > 0. \end{aligned} \end{equation*} $ |
Hence
$ \begin{equation*} (D_1D_2)^2 > ((ae^{-\zeta_1}-d)d_*)^2. \end{equation*} $ |
$ \begin{equation*} (D_1D_2)^2+(E_1E_2)^2+(D_1E_2)^2+(D_2E_1)^2 \ge (D_1D_2)^2 > ((ae^{-\zeta_1}-d)d_*)^2. \end{equation*} $ |
By Eq (3.5), we get
$ \begin{equation*} \begin{aligned} (F_1)^2+(F_2)^2\ge (D_1D_2)^2 & > ((ae^{-\zeta_1}-d)d_*)^2\\ (bce^{-\zeta_2})^2 (\cos^2(v\tau_2)+\sin^2(v\tau_2))&\ge (D_1D_2)^2 > ((ae^{-\zeta_1}-d)d_*)^2\\ (bce^{-\zeta_2})^2\ge (D_1D_2)^2& > ((ae^{-\zeta_1}-d)d_*)^2. \end{aligned} \end{equation*} $ |
By assumption $ (ae^{-\zeta_1}-d)d_*+bce^{-\zeta_2} > 0 $, which is a contradiction, thus $ u $ must be negative real parts. This completes the proof of Lemma 3.3.
Before the details, we will present the notion from the literature [25]. We define
$ x_t\in\mathcal{C}([-\tau, 0], \mathbb{R}^2), $ |
by $ x_t(\theta) = x(t+\theta), \forall \theta\in[-\tau, 0] $. Consider a delay system
$ \begin{equation} \begin{aligned} x'(t) = f(x_t), \end{aligned} \end{equation} $ | (3.6) |
for which uniqueness of solutions is assumed, $ x(t, \psi) $ designates the solution of Eq (3.6) with initial condition $ x_0 = \psi $ $ (\psi\in\mathcal{C}) $.
A non-negative equilibrium $ v = (v_p, v_m)\in \mathbb{R}^2 $ of system (2.2) is said to be globally attractive if $ Y(t)\to v $ as $ t\to\infty $, for all admissible solutions $ Y(t) $ of system (2.2). We say that $ v $ is globally asymptotically stable if it is stable and globally attractive.
System (2.2) is written as Eq (3.6),
$ \begin{equation} \begin{aligned} f_1(\psi) = &\psi_1(0)[-d-b_1\psi_1(0)]+ae^{-\zeta_1}\psi_1(-\tau_1)+b\psi_2(0), \\ f_2(\psi) = &\psi_2(0)[-d_*-b_2\psi_2(0)]+ce^{-\zeta_2}\psi_1(-\tau_2). \end{aligned} \end{equation} $ | (3.7) |
Observe that system (2.2) is cooperative, i.e., $ Df_i(\psi)\varphi\ge 0 $, for all $ \psi, \varphi\in X^+ $ with $ \varphi_i(0) = 0, i = 1, 2 $. This implies that $ f $ satisfies quasi-monotonicity condition [26,p. 78]. Typically, in population dynamics the stability of equilibria is closely related to the algebraic properties of some kinds of competition matrix of the community. Denote
$ \begin{equation*} A = \begin{pmatrix} ae^{-\zeta_1}-d&0\\ 0&-d_* \end{pmatrix}, \quad D = \begin{pmatrix} 0&b\\ ce^{-\zeta_2}&0 \end{pmatrix}. \end{equation*} $ |
For convenience, we shall refer to $ N = A+D $ as the (linear) community matrix:
$ \begin{equation} N = \begin{pmatrix} ae^{-\zeta_1}-d&b\\ ce^{-\zeta_2}&-d_* \end{pmatrix}. \end{equation} $ | (3.8) |
Since $ D\ge 0 $, the matrix $ N $ in Eq (3.8) is called cooperative. If $ D $ is irreducible, then the matrix $ N $ in Eq (3.8) is also irreducible; in this case, system (2.2) is called an irreducible system [26,p. 88], and the semiflow $ \psi\mapsto Y_t(\psi) $ is eventually strongly monotone. $ f = (f_1, f_2)^T:\mathbb{R}^2 \to \mathbb{R}^2 $ is strictly sublinear, i.e., for any $ P\gg 0, M\gg 0 $ and any $ \alpha\in(0, 1), $
$ \begin{equation*} \begin{aligned} f_1(\alpha P, \alpha M) = &\alpha P[-d-b_1\alpha P]+ae^{-\zeta_1}\alpha P(t-\tau_1)+b\alpha M\\ > & \alpha [P(-d-b_1P)+ae^{-\zeta_1} P(t-\tau_1)+b M] = \alpha f_1(P, M), \\ f_2(\alpha P, \alpha M) = &\alpha M[-d_*-b_2\alpha M]+ce^{-\zeta_2}\alpha P(t-\tau_2)\\ > &\alpha [M(-d_*-b_2M)+ce^{-\zeta_2}P(t-\tau_2)] = \alpha f_2(P, M). \end{aligned} \end{equation*} $ |
Cooperative $ DDEs $ satisfying these sublinearity conditions have significant properties [30,Proposition 4.3].
Recall that the stability modulus of square matrix $ N $ in Eq (3.8), denoted by $ s(N) $, is defined by $ s(N) = \max \{Re \lambda:\mbox{ $\lambda$ is an eigenvalue of $N$ }\rbrace $. If the matrix $ N $ in Eq (3.8) has nonnegative off diagonal elements and is irreducible, then $ s(N) $ is a simple eigenvalue of the matrix $ N $ with a (componentwise) positive eigenvector (see, e.g., [31,Theorem A.5]).
The matrix $ N $ in Eq (3.8) is
$ \begin{equation*} N = \begin{pmatrix} ae^{-\zeta_1}-d&b\\ ce^{-\zeta_2}&-d_* \end{pmatrix}, \end{equation*} $ |
then we can easily get the following:
$ s(N) > 0 $ if and only if $ (ae^{-\zeta_1}-d)d_*+bce^{-\zeta_2} > 0 $ and $ s(N) < 0 $ if and only if $ (ae^{-\zeta_1}-d)d_*+bce^{-\zeta_2} < 0 $.
Definition 3.4. [32] A square matrix $ A = [a_{ij}] $ with non-positive off diagonal entries, i.e., $ a_{ij}\le 0 $ for all $ i\neq j $, is said to be an M-matrix if all the eigenvalues of $ A $ have a non-negative real part, or equivalently, if all its principal minors are non-negative, and $ A $ is said to be a non singular M-matrix if all the eigenvalues of $ A $ have positive real part, or, equivalently, if all its principal minors are positive.
Theorem 3.5. Suppose that $ (ae^{-\zeta_1}-d)d_*+bce^{-\zeta_2} < 0 $, then the equilibrium $ E_0 $ of system (2.2) is globally asymptotically stable.
Proof. Let $ P(t, l) $, $ M(t, k) $ be the solutions of system (2.2) with $ P(0+\theta, l) = l $, $ M(0+\theta, k) = k $ for $ \theta\in[-\tau, 0] $. Note that $ f_1(l) = l[-d+b+ae^{-\zeta_1}-b_1l] < 0 $ for $ l > 0 $ sufficiently large and $ f_2(k) = k[-d_*+ce^{-\zeta_2}-b_2k] < 0 $ for $ k > 0 $ sufficiently large. Hence we can easily conclude that all admissible solutions of system (2.2) are bounded [26,Corollary 5.2.2]. We have $ s(N) < 0 $ if and only if $ (ae^{-\zeta_1}-d)d_*+bce^{-\zeta_2} < 0 $. By the assumption $ (ae^{-\zeta_1}-d)d_*+bce^{-\zeta_2} < 0 $, we observed that it is equivalent to having $ -N $ a non singular M-matrix. Since matrix $ -N $ is a non singular M-matrix, there exists the equilibrium $ v = (v_p, v_m)\in \mathbb{R}^2, v > 0, $ such that $ Nv < 0 $, hence we get
$ \begin{equation} \begin{aligned} ae^{-\zeta_1}v_p-dv_p+bv_m < 0, \\ ce^{-\zeta_2}v_p-d_*v_m < 0. \end{aligned} \end{equation} $ | (3.9) |
Let $ P(t)\ge 0 $, $ M(t)\ge 0 $ be solutions of system (2.2). Denote $ y_p(t) = \frac{P(t)}{v_p} $ and $ y_m(t) = \frac{M(t)}{v_m} $, thus system (2.2) takes the form as
$ \begin{equation} \begin{aligned} y_p'(t) = &y_p(t)[-d-b_1y_p(t)v_p]+ae^{-\zeta_1}y_p(t-\tau_1)+\frac{bv_m}{v_p}y_m(t), \\ y_m'(t) = &y_m(t)[-d_*-b_2y_m(t)v_m]+ce^{-\zeta_2}\frac{v_p}{v_m}y_p(t-\tau_2). \end{aligned} \end{equation} $ | (3.10) |
It suffices to prove that $ (L_p, L_m): = \limsup_{t\to\infty} (y_p(t), y_m(t)) = (0, 0) $. Let $ L_p: = \limsup\lbrace{y_p(t)}\rbrace $, $ L_m: = \limsup\lbrace{y_m(t)}\rbrace $, $ \tilde{L}: = \max\lbrace{L_p, L_m}\rbrace $ and suppose that $ \tilde{L} > 0 $. From Eq (3.9), we can choose $ \varepsilon > 0 $ such that
$ \begin{equation*} \begin{aligned} \tilde{L}[-d-b_1\tilde{L}v_p+ae^{-\zeta_1}+\frac{bv_m}{v_p}]+\varepsilon[ae^{-\zeta_1}+\frac{bv_m}{v_p}] = :\gamma_p < 0, \\ \tilde{L}[-d_*-b_2\tilde{L}v_m+ce^{-\zeta_2}\frac{v_p}{v_m}]+\varepsilon[ce^{-\zeta_2}\frac{v_p}{v_m}] = :\gamma_m < 0. \end{aligned} \end{equation*} $ |
Let $ T > 0 $ be such that $ y_p(t)\le \tilde{L}+\varepsilon $, $ y_m(t)\le \tilde{L}+\varepsilon $ for all $ t > T-\tau $ and the cases of $ y_p(t) $ and $ y_m(t) $ are separated as eventually monotone and not eventually monotone. By [26,Proposition 5.4.2], if $ y_p(t) $ and $ y_m(t) $ are eventually monotone, then $ y_p(t)\to \tilde{L} $ and $ y_m(t)\to \tilde{L} $ as $ t\to \infty $ for $ t\ge T $ and we obtain
$ \begin{equation} \begin{aligned} y'_p(t)&\le y_p(t)[-d-b_1y_p(t)v_p]+ae^{-\zeta_1}(\tilde{L}+\varepsilon)+(\tilde{L}+\varepsilon)\frac{bv_m}{v_p}\to \gamma_p, \\ y'_m(t)&\le y_m(t)[-d_*-b_2y_m(t)v_m]+ce^{-\zeta_2}\frac{v_p}{v_m}(\tilde{L}+\varepsilon)\to \gamma_m \text{ as } t\to\infty. \end{aligned} \end{equation} $ | (3.11) |
Since $ \gamma_p < 0 $ and $ \gamma_m < 0 $, these imply that $ \lim_{t\to \infty} (y_p(t), y_m(t)) = -\infty $, which is impossible. By using the similar argument of Aiello and Freedman [18,Theorem 2], if $ y_p(t) $ and $ y_m(t) $ are not eventually monotone, there is a sequence $ t_n\to\infty $ such that $ y_p(t_n)\to \tilde{L}, $ $ y'_p(t_n) = 0 $ and $ y_m(t_n)\to \tilde{L}, $ $ y'_m(t_n) = 0 $. We obtain (3.11) with $ t $ replaced by $ t_n $, again a contradiction. This proves $ \lim_{t\to \infty} (y_p(t), y_m(t)) = (0, 0) $. Using Lemma 3.1, we complete the proof of Theorem 3.5.
Theorem 3.6. Suppose that $ ae^{-\zeta_1}-d > 0 $ and $ c = 0 $, then the equilibrium $ E_1 $ of system (2.2) is globally asymptotically stable.
Proof. If $ c = 0 $, the second equation of system (2.2) becomes
$ \begin{equation} M'(t) = -d_*M-b_2M^2, \end{equation} $ | (3.12) |
For the independent subsystem (3.12), it is obvious that $ \lim_{t\to \infty} M(t) = 0 $.
Then the first equation of system (2.2) becomes
$ \begin{equation} P'(t) = ae^{-\zeta_1}P(t-\tau_1)-dP(t)-b_1P^2(t). \end{equation} $ | (3.13) |
Let $ \varepsilon > 0 $ be sufficiently small and $ L > 0 $ be sufficiently large such that $ \varepsilon\le P(t)\le L $, $ t\in[-\tau, 0] $, and
$ ae^{-\zeta_1}\varepsilon-d\varepsilon-b_1\varepsilon^2 > 0, \quad ae^{-\zeta_1}L-dL-b_1L^2 < 0. $ |
Let $ P_\varepsilon(t) $ and $ P_L(t) $ be the solutions of Eq (3.13) with $ P_\varepsilon(t) = \varepsilon $ and $ P_L(t) = L $ for $ t\in[-\tau, 0] $. From the monotone properties of the equation [26], the function $ P_\varepsilon(t) $ is increasing and $ P_L(t) $ is decreasing for $ t\ge 0 $ and
$ P_\varepsilon(t)\le P(t)\le P_L(t), t\ge 0. $ |
It therefore follows that
$ \begin{equation*} \begin{aligned} \frac{ae^{-\zeta_1}-d}{b_1} = \lim\limits_{t\to\infty} P(t)\le \lim\limits_{t\to\infty} P_L(t) = \frac{ae^{-\zeta_1}-d}{b_1} \end{aligned} \end{equation*} $ |
because the only equilibrium of the equation between $ \varepsilon $ and $ L $ is $ \frac{ae^{-\zeta_1}-d}{b_1} $. Using Lemma 3.2, we complete the proof of Theorem 3.6.
Lemma 3.7. Suppose there is a positive equilibrium $ (P^*, M^*) $ of system (2.2), and that $ (ae^{-\zeta_1}-d)d_*+bce^{-\zeta_2} > 0 $ and $ c\neq 0 $. Then all solutions $ P(t, \psi_1) $, $ M(t, \psi_2) $ of system (2.2) with $ \psi_i\in X_0^+, i = 1, 2 $ satisfy $ \liminf_{t\to\infty}(P(t, \psi_1), M(t, \psi_2))\ge (P^*, M^*) $.
Proof. For $ (P^*, M^*) $ an equilibrium of system (2.2), we have
$ \begin{equation} \begin{aligned} ae^{-\zeta_1}-d+b\frac{M^*}{P^*} = b_1P^* > 0, \\ -d_*+ce^{-\zeta_2}\frac{P^*}{M^*} = b_2M^* > 0. \end{aligned} \end{equation} $ | (3.14) |
Denote $ \bar{P}(t) = \frac{P(t)}{P^*} $ and $ \bar{M}(t) = \frac{M(t)}{M^*} $ in system (2.2), and dropping the bar for simplicity, we get
$ \begin{equation} \begin{aligned} P'(t) = &P(t)[-d-b_1P(t)P^*]+ae^{-\zeta_1}P(t-\tau_1)+\frac{bM^*}{P^*}M(t), \\ M'(t) = &M(t)[-d_*-b_2M(t)M^*]+ce^{-\zeta_2}\frac{P^*}{M^*}P(t-\tau_2). \end{aligned} \end{equation} $ | (3.15) |
For the solutions $ P(t) = P(t, \psi_1) $ and $ M(t) = M(t, \psi_2) $ of Eq (3.15) with $ \psi_i\in X_0^+ $ for $ i = 1, 2 $, we first claim that $ (l_p, l_m): = \liminf_{t\to\infty}(P(t), M(t)) > (0, 0) $. Otherwise, there exist $ \delta\in(0, 1) $ and $ t_0 > \tau $ such that $ \tilde{l} = \min\lbrace{(P(t), M(t)):t\in[0, t_0]}\rbrace $ and $ \tilde{l} < \delta $. By using Eq (3.14),
$ \begin{equation*} \begin{aligned} P'(t_0)& = P(t_0)[-d-b_1P(t_0)P^*]+ae^{-\zeta_1}P(t_0-\tau_1)+\frac{bM^*}{P^*}M(t_0)\\ &\ge \tilde{l}[-d-b_1\tilde{l}P^*]+ae^{-\zeta_1}\tilde{l}+\frac{bM^*}{P^*}\tilde{l} \\ & = \tilde{l}(b_1P^*-b_1\tilde{l}P^*) = \tilde{l}b_1P^*(1-\tilde{l}) > 0, \\ M'(t_0)& = M(t_0)[-d_*-b_2M(t_0)M^*]+ce^{-\zeta_2}\frac{P^*}{M^*}P(t_0-\tau_2)\\ &\ge \tilde{l}[-d_*-b_2\tilde{l}M^*]+ce^{-\zeta_2}\frac{P^*}{M^*}\tilde{l} \\ & = \tilde{l}(b_2M^*-b_2\tilde{l}M^*) = \tilde{l}b_2M^*(1-\tilde{l}) > 0. \end{aligned} \end{equation*} $ |
But these are not possible. Since the definition of $ t_0 $, $ P'(t_0)\le 0 $ and $ M'(t_0)\le 0 $.
Next we prove that $ (l_p, l_m)\ge(1, 1) $. Choose $ \tilde{l} = \min \lbrace{l_p, l_m}\rbrace $ and suppose that $ \tilde{l} < 1 $. Let $ T > 0 $ and $ \varepsilon > 0 $ be chosen so that $ P(t)\ge \tilde{l}-\varepsilon $ and $ M(t)\ge \tilde{l}-\varepsilon $ for all $ t > T-\tau $.
$ \begin{equation*} \begin{aligned} \tilde{l}b_1P^*(1-\tilde{l})-\varepsilon[ae^{-\zeta_1}\tilde{l}+\frac{bM^*}{P^*}] = :n_p > 0, \\ \tilde{l}b_2M^*(1-\tilde{l})-\varepsilon[ce^{-\zeta_2}\frac{P^*}{M^*}] = :n_m > 0. \end{aligned} \end{equation*} $ |
By [26,Proposition 5.4.2], if $ P(t) $ and $ M(t) $ are eventually monotone, then $ P(t)\to \tilde{l} $ and $ M(t)\to \tilde{l} $ and for $ t\ge T $, we have
$ \begin{equation*} \begin{aligned} P'(t)&\ge P(t)[-d-b_1P(t)P^*]+ae^{-\zeta_1}(\tilde{l}-\varepsilon)+(\tilde{l}-\varepsilon)\frac{bM^*}{P^*}\to n_p, \\ M'(t)&\ge M(t)[-d_*-b_2M(t)M^*]+(\tilde{l}-\varepsilon)ce^{-\zeta_2}\frac{P^*}{M^*} \to n_m \text{ as } t\to\infty, \end{aligned} \end{equation*} $ |
leading to $ P(t)\to \infty $ and $ M(t)\to \infty $ as $ t\to \infty $, contradicting $ \tilde{l} < 1 $. By using the similar argument of Aiello and Freedman [18,Theorem 2], if $ P(t) $ and $ M(t) $ are not eventually monotone, there is a sequence $ t_n\to \infty $ such that $ P(t_n)\to \tilde{l}, $ $ P'(t_n) = 0 $ and $ M(t_n)\to \tilde{l}, $ $ M'(t_n) = 0 $. For $ t_n\ge T $, we obtain the above inequalities $ t_n $ instead of $ t $, which yield that $ 0 = P'(t_n)\ge n_p $ and $ 0 = M'(t_n)\ge n_m $, again contradicting $ \tilde{l} < 1 $. This proves that $ \tilde{l}\ge 1 $.
Theorem 3.8. Suppose that $ (ae^{-\zeta_1}-d)d_*+bce^{-\zeta_2} > 0 $ and $ c\neq 0 $, then the equilibrium $ E^* $ of system (2.2) is globally asymptotically stable.
Proof. For $ (P^*, M^*) $ of system (2.2), after the changes $ P(t)\mapsto \frac{P(t)}{P^*} $ and $ M(t)\mapsto \frac{M(t)}{M^*} $, consider system (3.15) with positive equilibrium $ (1, 1)\in \mathbb{R}^2 $. In view of Lemmas 3.3 and 3.7, we only need to prove that $ (L_p, L_m): = \limsup_{t\to \infty}(P(t), M(t))\le (1, 1) $ and any positive solution $ P(t) $, $ M(t) $ of Eq (3.15).
For the sake of contradiction, suppose that $ \tilde{L} = \max \lbrace{L_p, L_m}\rbrace > 1 $. Choose $ \varepsilon > 0 $ and $ t > \tau $, such that $ P(t)\le \tilde{L}+\varepsilon $ and $ M(t)\le \tilde{L}+\varepsilon $ for all $ t > T-\tau $ and
$ \begin{equation*} \begin{aligned} \tilde{L}b_1P^*(1-\tilde{L})+\varepsilon[ae^{-\zeta_1}\tilde{L}+\frac{bM^*}{P^*}] = :&N_p < 0, \\ \tilde{L}b_2M^*(1-\tilde{L})+\varepsilon[ce^{-\zeta_2}\frac{P^*}{M^*}] = :& N_m < 0. \end{aligned} \end{equation*} $ |
Separating the cases of $ P(t) $ and $ M(t) $ eventually monotone and not eventually monotone, and reasoning as in the proofs of Theorem 3.5 and Lemma 3.7, we obtain a contradiction, thus $ \tilde{L}\le 1 $. Finally we get $ \lim_{t\to \infty}(P(t), M(t)) = (P^*, M^*) $. Using Lemma 3.3, we complete the proof of Theorem 3.8.
Remark 1. Note that when $ \tau_1 = \tau_2 = 0 $, system (2.2) becomes system (1.1). Theorems 4–6 in [11] are the corresponding results of Theorems 3.5, 3.6 and 3.8 for system (2.2), respectively. Our main results not only extend the results in [11] but also generalize the related results into the stage-structured system with two delays. But the proof methods of our results are quite different to those in [11].
In this section, we numerically simulate the dynamics of system (2.2) for a range of parameters which are the same as those in [11]. In this paper, we add the values of two delays $ \tau_1 $ and $ \tau_2 $ from [10,33]. The parameters are given in Table 1.
Parameter | Ranges | Ref. | Unit | data 1 | data 2 |
$ \alpha(T) $ | $ 0.03\sim0.15^{a} $ | [14] | ind $ \cdot $ d$ ^{-1} $ $ P^{-1} $ | 0.12 | 0.15 |
$ \beta(T) $ | $ 0.065\sim0.139^{a} $ | [14] | ind $ \cdot $ d$ ^{-1} $time$ ^{-1} $ $ P^{-1} $ | 0.108 | 0.122 |
$ \gamma $ | $ 19\sim178^{a} $ | [33,34] | ind $ \cdot $ d$ ^{-1} $ $ M^{-1} $ | 100 | 170 |
$ s_1 $ | $ 0.001\sim0.3^{b} $ | [34,35] | no unit | 0.008 | 0.01 |
$ s_2 $ | $ 0.01\sim0.8^{b} $ | [34] | no unit | 0.2 | 0.8 |
$ n $ | $ 1\sim2 $ | [14] | times | 1 | 1 |
$ d_1 $ | $ 0\sim0.028^{a, b} $ | [6,14] | d$ ^{-1} $ | 0.0001 | 0.0001 |
$ d_2 $ | $ 0.0001\sim0.3^{b} $ | [34] | d$ ^{-1} $ | 0.0001 | 0.0001 |
$ d_3 $ | $ 0.004\sim0.02^{a} $ | [34,36] | d$ ^{-1} $ | 0.006 | 0.004 |
$ d_4 $ | $ 0.0001\sim0.8^{b} $ | [1] | d$ ^{-1} $ | 0.0001 | 0.0001 |
$ b_1 $ | $ 0.00001\sim0.1^{b} $ | [3,37] | d$ ^{-1} $ind$ ^{-1} $ | 0.0012 | 0.0001 |
$ b_2 $ | $ 0\sim0.1^{b} $ | d$ ^{-1} $ind$ ^{-1} $ | 0.0001 | 0.0001 | |
$ \tau_1 $ | $ 30\sim120^{b} $ | [10,33] | d | 120 | 120 |
$ \tau_2 $ | $ 60\sim300^{b} $ | [10,33] | d | 90 | 150 |
Values signatured by $ ^{a} $ are from experimental data with unit innovation and those signatured by $ ^{b} $ are estimated from references.
The left figure of Figure 2 shows that the positive equilibrium $ E^* $ of system (2.2) is globally asymptotically stable under different initial values. The left figure and right figure of Figure 2 take the parameters data 1 and data 2, respectively. Figure 2 shows that the population sizes change with respect to environmental indices but do not depend on the initial values. The population explosion occurs even though the initial values $ (P = 0, M = 2) $ are small (see the right figure of Figure 2). The numbers of two stages in the right figure of Figure 2 are larger than those in the left figure of Figure 2 because the reproduction is high while the destructions and competitions are low in the right figure of Figure 2. The trajectories of the right figure of Figure 2 finally tend towards a higher population level up to 10–15 times than the trajectories in the left figure of Figure 2 (in the corresponding Figure 3 of [11], the populations of Figure 3(b) is higher 30–50 times than Figure 3(a)) although the initial values $ (0, 2) $ are of equal values.
Based on data 2, Figure 3 illustrates how time delays affect the population dynamics. In the left figure of Figure 3, we fix the delay $ \tau_2 $ as the best fit value and increase the delay $ \tau_1\in [30,120] $. We find that the populations are slightly fallen over the longer period $ \tau_1 $ (see the left figure of Figure 3). This is because of the lack of needed temperature and resources and so the asexual reproduction period is long, and the results of the population are low. When we fix the delay $ \tau_1 $ and change the delay $ \tau_2 $ from $ 60 $ to $ 300 $, the populations are significantly decreased over the longer period $ \tau_2 $ (see the right figure of Figure 3). Overall, Figure 3 can be seen that the peaks of population abundance occur at the small $ \tau_1 $ and $ \tau_2 $ while the longer maturation periods may be responsible for the lower populations.
Figure 4 depicts the effect of temperature $ T\in [7,36] $ on the populations. Temperature is the impact factor that affects the asexual reproduction and strobilation of the jellyfish. In [11], Xie et al. presented that
$ \alpha(T) = \frac{1.9272}{T^3-30.3904T^2+294.7234T-871.29}+0.0378, $ |
$ \beta(T) = 0.1430 \text{ exp } \lbrace{-(\frac{T-16.8108}{10.5302})^2}\rbrace. $ |
In Figure 4, the numbers of polyp reach a peak at $ 12.5\; ^{\circ}C $, which correlates with the maximum budding rate of experimental data [14] and then gradually declined over the high temperature. From $ 12.5\; ^{\circ}C $ to $ 16.8\; ^{\circ}C $ is the maximum level of the number of medusae which is different from the experimental result $ 15\; ^{\circ}C $ [14]. Figure 4 reveals that an appropriate increase of temperature might cause a large increase in the number of populations but the rise of temperatures would result in the fewer populations. Comparing the corresponding Figure (4d) in [11] with the right figure of Figure 4 in this paper, we find out that we can exactly see the peak populations due to the stage structure and can exactly know the effects of temperature on the population dynamics because the temperature is considered up to $ 36\; ^{\circ}C $ in this paper.
In this paper, we propose and analyze a delayed jellyfish model with stage structure, which is an extension of ODE model studied by Xie et al. in [11]. We have investigated how the phenomena of budding and strobilation influence the population dynamics of the jellyfish population. $ \tau_1 $ stands the time needed from the stage of the young polyp to the developed polyp and $ \tau_2 $ stands the time taken from the mature polyp to ephyra (incipient medusa). We have developed the systematic analysis of the model in both theoretical and numerical ways.
We have proved the global stability of the equilibria under suitable conditions. Our results not only extend but also improve some related results of literature [11]. Our Theorems 3.5, 3.6 and 3.8 straightly extend the corresponding Theorems 4–6 in [11], respectively. Comparing the corresponding Theorems 4–6 in [11] for the ODE system (1.1) with Theorem 3.5, 3.6 and 3.8 for system (2.2), we find out that there are two extra terms $ e^{-d\tau_1} $ and $ e^{-d_*\tau_2} $ in our permanence and extinction criteria, i.e., the surviving probability of each immature population to develop into mature, which obtains due to the stage structure. From our results, we find that the jellyfish population will be extinct in the large immature mortality rate $ d, d_* $ or the long maturation $ \tau_1, \tau_2 $. Thus we may suggest that the proper increases of $ d\tau_1 $ and $ d_*\tau_2 $ have a negative effect of jellyfish population.
Biologically, our results suggest that (i) jellyfish species go extinct if the survival rate of polyp during cloning and the survival rate of the incipient medusa during strobilation are less than their death rates; (ii) polyps will continue and there is no complement from polyp to medusa if the survival rate of polyp during cloning is larger than its death rate and the temperature is not enough to strobilate; (iii) both polyp and medusa will survive in a certain ideal environment and our result converges to the positive constant when the survival rate of polyp during cloning and the survival rate of the incipient medusa during strobilation are larger than their death rates.
Besides the above systematic theoretical results, we have performed the numerical simulations to support the theoretical results. Our numerical results suggest that the positive equilibrium is globally asymptotically stable under distinctive initial values and the population sizes don't deal with the initial values but they change with respect to environmental factors. In Figures 3 and 4, our results suggest that the abundance in population occurs at the smaller periods $ \tau_1 $ and $ \tau_2 $ whereas the longer periods $ \tau_1 $ and $ \tau_2 $ will lower the peak population of polyp and medusa. In addition to the problem due to increasing $ \tau_1 $ and $ \tau_2 $, the increase of temperatures might cause the outburst of the population dynamics. If there is much higher temperature, the population rate leads to decline. Since temperature has a great impact on jellyfish population, it is interesting for one to consider the populations under the relevance to temperature. We leave this interesting problem as our future work.
We would like to take this chance to thank the editor and the anonymous referees for their very valuable comments, which led to a significant improvement of our previous versions. The authors would like to thank Dr. Zhanwen Yang for his warm help on the numeric simulations. Z. Win, B. Tian and S. Liu are supported by the Natural Science Foundation of China (NSFC) (No. 11871179, 11771374, 91646106).
All authors declare no conflicts of interest in this paper.
[1] |
De Strooper B, Saftig P, Craessaerts K, et al. (1998) Deficiency of presenilin-1 inhibits the normal cleavage of amyloid precursor protein. Nature 391: 387-390. doi: 10.1038/34910
![]() |
[2] |
Jankowsky JL, Fadale DJ, Anderson J, et al. (2004) Mutant presenilins specifically elevate the levels of the 42 residue beta-amyloid peptide in vivo: evidence for augmentation of a 42-specific gamma secretase. Hum Mol Genet 13: 159-170. doi: 10.1093/hmg/ddh019
![]() |
[3] |
Clark RF, Hutton M, Fuldner M, et al. (1995) The structure of the presenilin 1 (S182) gene and identification of six novel mutations in early onset AD families. Nat Genet 11: 219-222. doi: 10.1038/ng1095-219
![]() |
[4] |
Hardy J, Selkoe DJ (2002) The amyloid hypothesis of Alzheimer's disease: progress and problems on the road to therapeutics. Science 297: 353-356. doi: 10.1126/science.1072994
![]() |
[5] |
Weksler ME, Gouras G, Relkin NR, et al. (2005) The immune system, amyloid-beta peptide, and Alzheimer's disease. Immunol Rev 205: 244-256. doi: 10.1111/j.0105-2896.2005.00264.x
![]() |
[6] |
Hass MR, Sato C, Kopan R, et al. (2009) Presenilin: RIP and beyond. Semin Cell Dev Biol 20: 201-210. doi: 10.1016/j.semcdb.2008.11.014
![]() |
[7] |
de Calignon A, Fox LM, Pitstick R, et al. (2010) Caspase activation precedes and leads to tangles. Nature 464: 1201-1204. doi: 10.1038/nature08890
![]() |
[8] |
Sleegers K, Lambert JC, Bertram L, et al. (2010) The pursuit of susceptibility genes for Alzheimer's disease: progress and prospects. Trends Genet 26: 84-93. doi: 10.1016/j.tig.2009.12.004
![]() |
[9] |
Bertram L, Tanzi RE (2008) Thirty years of Alzheimer's disease genetics: the implications of systematic meta-analyses. Nat Rev Neurosci 9: 768-778. doi: 10.1038/nrn2494
![]() |
[10] |
Selkoe DJ, Hardy J (2016) The amyloid hypothesis of Alzheimer's disease at 25 years. EMBO Mol Med2016: e201606210. doi: 10.15252/emmm.201606210
![]() |
[11] |
Jones L, Holmans PA, Hamshere ML, et al. (2010) Genetic evidence implicates the immune system and cholesterol metabolism in the aetiology of Alzheimer's disease. PLoS One 5: e13950. doi: 10.1371/journal.pone.0013950
![]() |
[12] |
Porcellini E, Carbone I, Ianni M, et al. (2010) Alzheimer's disease gene signature says: beware of brain viral infections. Immun Ageing 7: 16. doi: 10.1186/1742-4933-7-16
![]() |
[13] |
Wozniak MA, Itzhaki RF, Shipley SJ, et al. (2007) Herpes simplex virus infection causes cellular beta-amyloid accumulation and secretase upregulation. Neurosci Lett 429: 95-100. doi: 10.1016/j.neulet.2007.09.077
![]() |
[14] |
Wozniak MA, Mee AP, Itzhaki RF (2009) Herpes simplex virus type 1 DNA is located within Alzheimer's disease amyloid plaques. J Pathol 217: 131-138. doi: 10.1002/path.2449
![]() |
[15] |
Santana S, Recuero M, Bullido MJ, et al. (2012) Herpes simplex virus type I induces the accumulation of intracellular beta-amyloid in autophagic compartments and the inhibition of the non-amyloidogenic pathway in human neuroblastoma cells. Neurobiol Aging 33: 430.e419-433. doi: 10.1016/j.neurobiolaging.2010.12.010
![]() |
[16] |
Seshadri S, Fitzpatrick AL, Ikram MA, et al. (2010) Genome-wide analysis of genetic loci associated with Alzheimer disease. JAMA 303: 1832-1840. doi: 10.1001/jama.2010.574
![]() |
[17] |
Carter CJ (2010) Alzheimer's Disease: A Pathogenetic Autoimmune Disorder Caused by Herpes Simplex in a Gene-Dependent Manner. Int J Alzheimers Dis 2010: 140539. doi: 10.4061/2010/140539
![]() |
[18] |
Tolia A, De Strooper B (2009) Structure and function of γ-secretase. Semin Cell Dev Biol 20: 211-218. doi: 10.1016/j.semcdb.2008.10.007
![]() |
[19] |
Edbauer D, Winkler E, Regula JT, et al. (2003) Reconstitution of gamma-secretase activity. Nat Cell Biol 5: 486-488. doi: 10.1038/ncb960
![]() |
[20] |
Kimberly WT, LaVoie MJ, Ostaszewski BL, et al. (2003) Gamma-secretase is a membrane protein complex comprised of presenilin, nicastrin, Aph-1, and Pen-2. Proc Natl Acad Sci U S A 100: 6382-6387. doi: 10.1073/pnas.1037392100
![]() |
[21] |
Takasugi N, Tomita T, Hayashi I, et al. (2003) The role of presenilin cofactors in the gamma-secretase complex. Nature 422: 438-441. doi: 10.1038/nature01506
![]() |
[22] |
Fraering PC, Ye W, Strub JM, et al. (2004) Purification and characterization of the human gamma-secretase complex. Biochemistry 43: 9774-9789. doi: 10.1021/bi0494976
![]() |
[23] |
Winkler E, Hobson S, Fukumori A, et al. (2009) Purification, pharmacological modulation, and biochemical characterization of interactors of endogenous human gamma-secretase. Biochemistry 48: 1183-1197. doi: 10.1021/bi801204g
![]() |
[24] |
Sato T, Diehl TS, Narayanan S, et al. (2007) Active gamma-secretase complexes contain only one of each component. J Biol Chem 282: 33985-33993. doi: 10.1074/jbc.M705248200
![]() |
[25] |
Hebert SS, Serneels L, Dejaegere T, et al. (2004) Coordinated and widespread expression of gamma-secretase in vivo: evidence for size and molecular heterogeneity. Neurobiol Dis 17: 260-272. doi: 10.1016/j.nbd.2004.08.002
![]() |
[26] |
Shirotani K, Edbauer D, Prokop S, et al. (2004) Identification of distinct gamma-secretase complexes with different APH-1 variants. J Biol Chem 279: 41340-41345. doi: 10.1074/jbc.M405768200
![]() |
[27] |
Beel AJ, Sanders CR (2008) Substrate specificity of gamma-secretase and other intramembrane proteases. Cell Mol Life Sci 65: 1311-1334. doi: 10.1007/s00018-008-7462-2
![]() |
[28] |
Coen K, Annaert W (2010) Presenilins: how much more than gamma-secretase?! Biochem Soc Trans 38: 1474-1478. doi: 10.1042/BST0381474
![]() |
[29] |
McCarthy JV, Twomey C, Wujek P (2009) Presenilin-dependent regulated intramembrane proteolysis and γ-secretase activity. Cell Mol Life Sci 66: 1534-1555. doi: 10.1007/s00018-009-8435-9
![]() |
[30] |
Wolfe MS (2009) Intramembrane-cleaving Proteases. J Biol Chem 284: 13969-13973. doi: 10.1074/jbc.R800039200
![]() |
[31] |
Brown MS, Ye J, Rawson RB, et al. (2000) Regulated intramembrane proteolysis: a control mechanism conserved from bacteria to humans. Cell 100: 391-398. doi: 10.1016/S0092-8674(00)80675-3
![]() |
[32] |
McCarthy JV, Twomey C, Wujek P (2009) Presenilin-dependent regulated intramembrane proteolysis and gamma-secretase activity. Cell Mol Life Sci 66: 1534-1555. doi: 10.1007/s00018-009-8435-9
![]() |
[33] |
Chhibber-Goel J, Coleman-Vaughan C, Agrawal V, et al. (2016) gamma-Secretase Activity Is Required for Regulated Intramembrane Proteolysis of Tumor Necrosis Factor (TNF) Receptor 1 and TNF-mediated Pro-apoptotic Signalling. J Biol Chem 291: 5971-5985. doi: 10.1074/jbc.M115.679076
![]() |
[34] | Elzinga BM, Twomey C, Powell JC, et al. (2009) Interleukin-1 receptor type 1 is a substrate for gamma-secretase-dependent regulated intramembrane proteolysis. J Biol Chem 284: 1394-1409. |
[35] |
Kuhn PH, Marjaux E, Imhof A, et al. (2007) Regulated intramembrane proteolysis of the interleukin-1 receptor II by alpha-, beta-, and gamma-secretase. J Biol Chem 282: 11982-11995. doi: 10.1074/jbc.M700356200
![]() |
[36] |
Chalaris A, Gewiese J, Paliga K, et al. (2010) ADAM17-mediated shedding of the IL6R induces cleavage of the membrane stub by gamma-secretase. Biochim Biophys Acta 1803: 234-245. doi: 10.1016/j.bbamcr.2009.12.001
![]() |
[37] |
Schulte A, Schulz B, Andrzejewski MG, et al. (2007) Sequential processing of the transmembrane chemokines CX3CL1 and CXCL16 by alpha- and gamma-secretases. Biochem Biophys Res Commun 358: 233-240. doi: 10.1016/j.bbrc.2007.04.100
![]() |
[38] |
Donoviel DB, Hadjantonakis AK, Ikeda M, et al. (1999) Mice lacking both presenilin genes exhibit early embryonic patterning defects. Genes Dev 13: 2801-2810. doi: 10.1101/gad.13.21.2801
![]() |
[39] |
Xia X, Qian S, Soriano S, et al. (2001) Loss of presenilin 1 is associated with enhanced beta-catenin signalling and skin tumorigenesis. Proc Natl Acad Sci U S A 98: 10863-10868. doi: 10.1073/pnas.191284198
![]() |
[40] |
Guo Q, Fu W, Sopher BL, et al. (1999) Increased vulnerability of hippocampal neurons to excitotoxic necrosis in presenilin-1 mutant knock-in mice. Nat Med 5: 101-106. doi: 10.1038/4789
![]() |
[41] |
Feng R, Rampon C, Tang YP, et al. (2001) Deficient neurogenesis in forebrain-specific presenilin-1 knockout mice is associated with reduced clearance of hippocampal memory traces. Neuron 32: 911-926. doi: 10.1016/S0896-6273(01)00523-2
![]() |
[42] |
Wang R, Dineley KT, Sweatt JD, et al. (2004) Presenilin 1 familial Alzheimer's disease mutation leads to defective associative learning and impaired adult neurogenesis. Neuroscience 126: 305-312. doi: 10.1016/j.neuroscience.2004.03.048
![]() |
[43] |
Herreman A, Hartmann D, Annaert W, et al. (1999) Presenilin 2 deficiency causes a mild pulmonary phenotype and no changes in amyloid precursor protein processing but enhances the embryonic lethal phenotype of presenilin 1 deficiency. Proc Natl Acad Sci U S A 96: 11872-11877. doi: 10.1073/pnas.96.21.11872
![]() |
[44] |
Tournoy J, Bossuyt X, Snellinx A, et al. (2004) Partial loss of presenilins causes seborrheic keratosis and autoimmune disease in mice. Hum Mol Genet 13: 1321-1331. doi: 10.1093/hmg/ddh151
![]() |
[45] |
Maraver A, Tadokoro CE, Badura ML, et al. (2007) Effect of presenilins in the apoptosis of thymocytes and homeostasis of CD8+ T cells. Blood 110: 3218-3225. doi: 10.1182/blood-2007-01-070359
![]() |
[46] |
Laky K, Fowlkes BJ (2007) Presenilins regulate αβ T cell development by modulating TCR signalling. J Exp Med 204: 2115-2129. doi: 10.1084/jem.20070550
![]() |
[47] |
Yagi T, Giallourakis C, Mohanty S, et al. (2008) Defective signal transduction in B lymphocytes lacking presenilin proteins. Proc Natl Acad Sci U S A 105: 979-984. doi: 10.1073/pnas.0707755105
![]() |
[48] |
Beglopoulos V, Sun X, Saura CA, et al. (2004) Reduced beta-amyloid production and increased inflammatory responses in presenilin conditional knock-out mice. J Biol Chem 279: 46907-46914. doi: 10.1074/jbc.M409544200
![]() |
[49] |
Jayadev S, Case A, Eastman AJ, et al. (2010) Presenilin 2 Is the Predominant γ-Secretase in Microglia and Modulates Cytokine Release. PLoS ONE 5: e15743. doi: 10.1371/journal.pone.0015743
![]() |
[50] |
Glenn G, van der Geer P (2008) Toll-like receptors stimulate regulated intramembrane proteolysis of the CSF-1 receptor through Erk activation. FEBS Lett 582: 911-915. doi: 10.1016/j.febslet.2008.02.029
![]() |
[51] |
Garlind A, Brauner A, Hojeberg B, et al. (1999) Soluble interleukin-1 receptor type II levels are elevated in cerebrospinal fluid in Alzheimer's disease patients. Brain Res 826: 112-116. doi: 10.1016/S0006-8993(99)01092-6
![]() |
[52] |
Dinarello CA (2011) Interleukin-1 in the pathogenesis and treatment of inflammatory diseases. Blood 117: 3720-3732. doi: 10.1182/blood-2010-07-273417
![]() |
[53] |
Twomey C, Qian S, McCarthy JV (2009) TRAF6 promotes ubiquitination and regulated intramembrane proteolysis of IL-1R1. Biochem Biophys Res Commun 381: 418-423. doi: 10.1016/j.bbrc.2009.02.051
![]() |
[54] |
Powell JC, Twomey C, Jain R, et al. (2009) Association between Presenilin-1 and TRAF6 modulates regulated intramembrane proteolysis of the p75NTR neurotrophin receptor. J Neurochem 108: 216-230. doi: 10.1111/j.1471-4159.2008.05763.x
![]() |
[55] | Curtis BM, Widmer MB, deRoos P, et al. (1990) IL-1 and its receptor are translocated to the nucleus. J Immunol 144: 1295-1303. |
[56] |
Levine SJ (2008) Molecular Mechanisms of Soluble Cytokine Receptor Generation. J Biol Chem 283: 14177-14181. doi: 10.1074/jbc.R700052200
![]() |
[57] |
Frykman S, Hur JY, Franberg J, et al. (2010) Synaptic and endosomal localization of active gamma-secretase in rat brain. PLoS One 5: e8948. doi: 10.1371/journal.pone.0008948
![]() |
[58] | Meckler X, Checler F (2016) Presenilin 1 and Presenilin 2 Target gamma-Secretase Complexes to Distinct Cellular Compartments. J Biol ChemM115: 708297. |
[59] |
Doody RS, Raman R, Farlow M, et al. (2013) A phase 3 trial of semagacestat for treatment of Alzheimer's disease. N Engl J Med 369: 341-350. doi: 10.1056/NEJMoa1210951
![]() |
[60] |
Goodbourn S, Didcock L, Randall RE (2000) Interferons: cell signalling, immune modulation, antiviral response and virus countermeasures. J Gen Virol 81: 2341-2364. doi: 10.1099/0022-1317-81-10-2341
![]() |
[61] |
Saleh AZ, Fang AT, Arch AE, et al. (2004) Regulated proteolysis of the IFNaR2 subunit of the interferon-alpha receptor. Oncogene 23: 7076-7086. doi: 10.1038/sj.onc.1207955
![]() |
[62] |
Garton KJ, Gough PJ, Raines EW (2006) Emerging roles for ectodomain shedding in the regulation of inflammatory responses. J Leukoc Biol 79: 1105-1116. doi: 10.1189/jlb.0106038
![]() |
[63] |
Mambole A, Baruch D, Nusbaum P, et al. (2008) The cleavage of neutrophil leukosialin (CD43) by cathepsin G releases its extracellular domain and triggers its intramembrane proteolysis by presenilin/gamma-secretase. J Biol Chem 283: 23627-23635. doi: 10.1074/jbc.M710286200
![]() |
[64] |
Cui W, Ke JZ, Zhang Q, et al. (2006) The intracellular domain of CD44 promotes the fusion of macrophages. Blood 107: 796-805. doi: 10.1182/blood-2005-05-1902
![]() |
[65] |
Carey BW, Kim DY, Kovacs DM (2007) Presenilin/gamma-secretase and alpha-secretase-like peptidases cleave human MHC Class I proteins. Biochem J 401: 121-127. doi: 10.1042/BJ20060847
![]() |
[66] |
Bonifati DM, Kishore U (2007) Role of complement in neurodegeneration and neuroinflammation. Mol Immunol 44: 999-1010. doi: 10.1016/j.molimm.2006.03.007
![]() |
[67] |
Veerhuis R, Nielsen HM, Tenner AJ (2011) Complement in the brain. Mol Immunol 48: 1592-1603. doi: 10.1016/j.molimm.2011.04.003
![]() |
[68] |
Fonseca MI, Chu SH, Berci AM, et al. (2011) Contribution of complement activation pathways to neuropathology differs among mouse models of Alzheimer's disease. J Neuroinflammation 8: 4. doi: 10.1186/1742-2094-8-4
![]() |
[69] |
Webster SD, Galvan MD, Ferran E, et al. (2001) Antibody-Mediated Phagocytosis of the Amyloid β-Peptide in Microglia Is Differentially Modulated by C1q. J Immunol 166: 7496-7503. doi: 10.4049/jimmunol.166.12.7496
![]() |
[70] |
Fonseca MI, Zhou J, Botto M, et al. (2004) Absence of C1q leads to less neuropathology in transgenic mouse models of Alzheimer's disease. J Neurosci 24: 6457-6465. doi: 10.1523/JNEUROSCI.0901-04.2004
![]() |
[71] |
Fraser DA, Pisalyaput K, Tenner AJ (2010) C1q enhances microglial clearance of apoptotic neurons and neuronal blebs, and modulates subsequent inflammatory cytokine production. J Neurochem 112: 733-743. doi: 10.1111/j.1471-4159.2009.06494.x
![]() |
[72] |
Rogers J, Li R, Mastroeni D, et al. (2006) Peripheral clearance of amyloid beta peptide by complement C3-dependent adherence to erythrocytes. Neurobiol Aging 27: 1733-1739. doi: 10.1016/j.neurobiolaging.2005.09.043
![]() |
[73] |
Akiyama H, Barger S, Barnum S, et al. (2000) Inflammation and Alzheimer's disease. Neurobiol Aging 21: 383-421. doi: 10.1016/S0197-4580(00)00124-X
![]() |
[74] |
Spitzer P, Herrmann M, Klafki H-W, et al. (2010) Phagocytosis and LPS alter the maturation state of β-amyloid precursor protein and induce different Aβ peptide release signatures in human mononuclear phagocytes. J Neuroinflammation 7: 59-59. doi: 10.1186/1742-2094-7-59
![]() |
[75] |
Morgan D, Gordon MN, Tan J, et al. (2005) Dynamic complexity of the microglial activation response in transgenic models of amyloid deposition: implications for Alzheimer therapeutics. J Neuropathol Exp Neurol 64: 743-753. doi: 10.1097/01.jnen.0000178444.33972.e0
![]() |
[76] |
Streit WJ (2004) Microglia and Alzheimer's disease pathogenesis. J Neurosci Res 77: 1-8. doi: 10.1002/jnr.20093
![]() |
[77] | Trudler D, Farfara D, Frenkel D (2010) Toll-Like Receptors Expression and Signalling in Glia Cells in Neuro-Amyloidogenic Diseases: Towards Future Therapeutic Application. Mediators of Inflammation 2010: 12. |
[78] |
Kumar H, Kawai T, Akira S (2011) Pathogen recognition by the innate immune system. Int Rev Immunol 30: 16-34. doi: 10.3109/08830185.2010.529976
![]() |
[79] |
Carty M, Bowie AG (2011) Evaluating the role of Toll-like receptors in diseases of the central nervous system. Biochem Pharmacol 81: 825-837. doi: 10.1016/j.bcp.2011.01.003
![]() |
[80] |
Chen F, Hasegawa H, Schmitt-Ulms G, et al. (2006) TMP21 is a presenilin complex component that modulates [gamma]-secretase but not [epsiv]-secretase activity. Nature 440: 1208-1212. doi: 10.1038/nature04667
![]() |
[81] | Iribarren P, Chen K, Hu J, et al. (2005) CpG-containing oligodeoxynucleotide promotes microglial cell uptake of amyloid beta 1-42 peptide by up-regulating the expression of the G-protein- coupled receptor mFPR2. FASEBJ 19: 2032-2034. |
[82] |
Jin JJ, Kim HD, Maxwell JA, et al. (2008) Toll-like receptor 4-dependent upregulation of cytokines in a transgenic mouse model of Alzheimer's disease. J Neuroinflammation 5: 23. doi: 10.1186/1742-2094-5-23
![]() |
[83] |
Herber DL, Mercer M, Roth LM, et al. (2007) Microglial activation is required for Abeta clearance after intracranial injection of lipopolysaccharide in APP transgenic mice. J Neuroimmune Pharmacol 2: 222-231. doi: 10.1007/s11481-007-9069-z
![]() |
[84] |
DiCarlo G, Wilcock D, Henderson D, et al. (2001) Intrahippocampal LPS injections reduce Abeta load in APP+PS1 transgenic mice. Neurobiol Aging 22: 1007-1012. doi: 10.1016/S0197-4580(01)00292-5
![]() |
[85] |
Richard KL, Filali M, Prefontaine P, et al. (2008) Toll-like receptor 2 acts as a natural innate immune receptor to clear amyloid beta 1-42 and delay the cognitive decline in a mouse model of Alzheimer's disease. J Neurosci 28: 5784-5793. doi: 10.1523/JNEUROSCI.1146-08.2008
![]() |
[86] |
Jana M, Palencia CA, Pahan K (2008) Fibrillar amyloid-beta peptides activate microglia via TLR2: implications for Alzheimer's disease. J Immunol 181: 7254-7262. doi: 10.4049/jimmunol.181.10.7254
![]() |
[87] |
Frank S, Copanaki E, Burbach GJ, et al. (2009) Differential regulation of toll-like receptor mRNAs in amyloid plaque-associated brain tissue of aged APP23 transgenic mice. Neurosci Lett 453: 41-44. doi: 10.1016/j.neulet.2009.01.075
![]() |
[88] |
Michaud J-P, Richard KL, Rivest S (2011) MyD88-adaptor protein acts as a preventive mechanism for memory deficits in a mouse model of Alzheimer's disease. Mol Neurodegener 6: 5-5. doi: 10.1186/1750-1326-6-5
![]() |
[89] |
Cui JG, Li YY, Zhao Y, et al. (2010) Differential regulation of interleukin-1 receptor-associated kinase-1 (IRAK-1) and IRAK-2 by microRNA-146a and NF-kappaB in stressed human astroglial cells and in Alzheimer disease. J Biol Chem 285: 38951-38960. doi: 10.1074/jbc.M110.178848
![]() |
[90] |
Standridge JB (2006) Vicious cycles within the neuropathophysiologic mechanisms of Alzheimer's disease. Curr Alzheimer Res 3: 95-108. doi: 10.2174/156720506776383068
![]() |
[91] |
Lue LF, Walker DG, Rogers J (2001) Modeling microglial activation in Alzheimer's disease with human postmortem microglial cultures. Neurobiol Aging 22: 945-956. doi: 10.1016/S0197-4580(01)00311-6
![]() |
[92] | Yan Q, Zhang J, Liu H, et al. (2003) Anti-inflammatory drug therapy alters beta-amyloid processing and deposition in an animal model of Alzheimer's disease. J Neurosci 23: 7504-7509. |
[93] |
Yamamoto M, Kiyota T, Horiba M, et al. (2007) Interferon-γ and Tumor Necrosis Factor-α Regulate Amyloid-β Plaque Deposition and β-Secretase Expression in Swedish Mutant APP Transgenic Mice. Am J Pathol 170: 680-692. doi: 10.2353/ajpath.2007.060378
![]() |
[94] |
Kong Q, Peterson TS, Baker O, et al. (2009) Interleukin-1beta enhances nucleotide-induced and alpha-secretase-dependent amyloid precursor protein processing in rat primary cortical neurons via up-regulation of the P2Y(2) receptor. J Neurochem 109: 1300-1310. doi: 10.1111/j.1471-4159.2009.06048.x
![]() |
[95] |
Sheng JG, Zhu SG, Jones RA, et al. (2000) Interleukin-1 Promotes Expression and Phosphorylation of Neurofilament and tau Proteins in Vivo. Exp Neurol 163: 388-391. doi: 10.1006/exnr.2000.7393
![]() |
[96] |
Moses GS, Jensen MD, Lue LF, et al. (2006) Secretory PLA2-IIA: a new inflammatory factor for Alzheimer's disease. J Neuroinflammation 3: 28. doi: 10.1186/1742-2094-3-28
![]() |
[97] |
Rezai-Zadeh K, Gate D, Gowing G, et al. (2011) How to Get from Here to There: Macrophage Recruitment in Alzheimer's Disease. Curr Alzheimer Res 8: 156-163. doi: 10.2174/156720511795256017
![]() |
[98] |
Shaftel SS, Kyrkanides S, Olschowka JA, et al. (2007) Sustained hippocampal IL-1 beta overexpression mediates chronic neuroinflammation and ameliorates Alzheimer plaque pathology. J Clin Invest 117: 1595-1604. doi: 10.1172/JCI31450
![]() |
[99] |
Rogers J, Strohmeyer R, Kovelowski CJ, et al. (2002) Microglia and inflammatory mechanisms in the clearance of amyloid beta peptide. Glia 40: 260-269. doi: 10.1002/glia.10153
![]() |
[100] |
Jutras I, Laplante A, Boulais J, et al. (2005) Gamma-secretase is a functional component of phagosomes. J Biol Chem 280: 36310-36317. doi: 10.1074/jbc.M504069200
![]() |
[101] |
Kiyota T, Yamamoto M, Xiong H, et al. (2009) CCL2 Accelerates Microglia-Mediated Aβ Oligomer Formation and Progression of Neurocognitive Dysfunction. PLoS ONE 4: e6197. doi: 10.1371/journal.pone.0006197
![]() |
[102] |
El Khoury J, Toft M, Hickman SE, et al. (2007) Ccr2 deficiency impairs microglial accumulation and accelerates progression of Alzheimer-like disease. Nat Med 13: 432-438. doi: 10.1038/nm1555
![]() |
[103] |
De Strooper B, Chavez Gutierrez L (2015) Learning by failing: ideas and concepts to tackle gamma-secretases in Alzheimer's disease and beyond. Annu Rev Pharmacol Toxicol 55: 419-437. doi: 10.1146/annurev-pharmtox-010814-124309
![]() |
[104] | Beher D, Clarke EE, Wrigley JD, et al. (2004) Selected non-steroidal anti-inflammatory drugs and their derivatives target gamma-secretase at a novel site. Evidence for an allosteric mechanism. J Biol Chem 279: 43419-43426. |
[105] |
Kukar T, Golde TE (2008) Possible mechanisms of action of NSAIDs and related compounds that modulate gamma-secretase cleavage. Curr Top Med Chem 8: 47-53. doi: 10.2174/156802608783334042
![]() |
[106] |
Wong GT, Manfra D, Poulet FM, et al. (2004) Chronic treatment with the gamma-secretase inhibitor LY-411,575 inhibits beta-amyloid peptide production and alters lymphopoiesis and intestinal cell differentiation. J Biol Chem 279: 12876-12882. doi: 10.1074/jbc.M311652200
![]() |
[107] |
Jack C, Berezovska O, Wolfe MS, et al. (2001) Effect of PS1 deficiency and an APP gamma-secretase inhibitor on Notch1 signalling in primary mammalian neurons. Brain Res Mol Brain Res 87: 166-174. doi: 10.1016/S0169-328X(01)00010-9
![]() |
[108] |
Geling A, Steiner H, Willem M, et al. (2002) A gamma-secretase inhibitor blocks Notch signalling in vivo and causes a severe neurogenic phenotype in zebrafish. EMBO Rep 3: 688-694. doi: 10.1093/embo-reports/kvf124
![]() |
[109] | Micchelli CA, Esler WP, Kimberly WT, et al. (2003) Gamma-secretase/presenilin inhibitors for Alzheimer's disease phenocopy Notch mutations in Drosophila. FASEB J 17: 79-81. |
[110] |
Henley DB, Sundell KL, Sethuraman G, et al. (2014) Safety profile of semagacestat, a gamma-secretase inhibitor: IDENTITY trial findings. Curr Med Res Opin 30: 2021-2032. doi: 10.1185/03007995.2014.939167
![]() |
[111] |
Panza F, Frisardi V, Imbimbo BP, et al. (2010) REVIEW: gamma-Secretase inhibitors for the treatment of Alzheimer's disease: The current state. CNS Neurosci Ther 16: 272-284. doi: 10.1111/j.1755-5949.2010.00164.x
![]() |
[112] |
Frisardi V, Solfrizzi V, Imbimbo PB, et al. (2010) Towards disease-modifying treatment of Alzheimer's disease: drugs targeting beta-amyloid. Curr Alzheimer Res 7: 40-55. doi: 10.2174/156720510790274400
![]() |
[113] |
Bergmans BA, De Strooper B (2010) gamma-secretases: from cell biology to therapeutic strategies. Lancet Neurol 9: 215-226. doi: 10.1016/S1474-4422(09)70332-1
![]() |
[114] | Samson K (2010) NerveCenter: Phase III Alzheimer trial halted: Search for therapeutic biomarkers continues. Ann Neurol 68: A9-a12. |
[115] |
Gillman KW, Starrett JE, Jr., Parker MF, et al. (2010) Discovery and Evaluation of BMS-708163, a Potent, Selective and Orally Bioavailable gamma-Secretase Inhibitor. ACS Med Chem Lett 1: 120-124. doi: 10.1021/ml1000239
![]() |
[116] |
De Strooper B (2014) Lessons from a failed gamma-secretase Alzheimer trial. Cell 159: 721-726. doi: 10.1016/j.cell.2014.10.016
![]() |
[117] |
Cummings J (2010) What Can Be Inferred from the Interruption of the Semagacestat Trial for Treatment of Alzheimer's Disease? Biol Psychiatry 68: 876-878. doi: 10.1016/j.biopsych.2010.09.020
![]() |
[118] |
Qyang Y, Chambers SM, Wang P, et al. (2004) Myeloproliferative disease in mice with reduced presenilin gene dosage: effect of gamma-secretase blockage. Biochemistry 43: 5352-5359. doi: 10.1021/bi049826u
![]() |
[119] |
Imbimbo BP (2008) Therapeutic potential of gamma-secretase inhibitors and modulators. Curr Top Med Chem 8: 54-61. doi: 10.2174/156802608783334015
![]() |
[120] |
Pul R, Dodel R, Stangel M (2011) Antibody-based therapy in Alzheimer's disease. Expert Opin Biol Ther 11: 343-357. doi: 10.1517/14712598.2011.552884
![]() |
[121] |
Sarazin M, Dorothee G, de Souza LC, et al. (2013) Immunotherapy in Alzheimer's disease: do we have all the pieces of the puzzle? Biol Psychiatry 74: 329-332. doi: 10.1016/j.biopsych.2013.04.011
![]() |
[122] |
Brodaty H, Breteler MM, Dekosky ST, et al. (2011) The world of dementia beyond 2020. J Am Geriatr Soc 59: 923-927. doi: 10.1111/j.1532-5415.2011.03365.x
![]() |
[123] |
Yamada K, Yabuki C, Seubert P, et al. (2009) Abeta immunotherapy: intracerebral sequestration of Abeta by an anti-Abeta monoclonal antibody 266 with high affinity to soluble Abeta. J Neurosci 29: 11393-11398. doi: 10.1523/JNEUROSCI.2021-09.2009
![]() |
[124] |
Schenk D, Barbour R, Dunn W, et al. (1999) Immunization with amyloid-[beta] attenuates Alzheimer-disease-like pathology in the PDAPP mouse. Nature 400: 173-177. doi: 10.1038/22124
![]() |
[125] |
Lemere CA, Spooner ET, Leverone JF, et al. (2003) Amyloid-beta immunization in Alzheimer's disease transgenic mouse models and wildtype mice. Neurochem Res 28: 1017-1027. doi: 10.1023/A:1023203122036
![]() |
[126] |
Fu HJ, Liu B, Frost JL, et al. (2010) Amyloid-beta immunotherapy for Alzheimer's disease. CNS Neurol Disord Drug Targets 9: 197-206. doi: 10.2174/187152710791012017
![]() |
[127] | Robert R, Dolezal O, Waddington L, et al. (2009) Engineered antibody intervention strategies for Alzheimer's disease and related dementias by targeting amyloid and toxic oligomers. Protein Eng Des Sel 22: 199-208. |
[128] |
Marin-Argany M, Rivera-Hernandez G, Marti J, et al. (2011) An anti-Abeta (amyloid beta) single-chain variable fragment prevents amyloid fibril formation and cytotoxicity by withdrawing Abeta oligomers from the amyloid pathway. Biochem J 437: 25-34. doi: 10.1042/BJ20101712
![]() |
[129] |
Panza F, Frisardi V, Imbimbo BP, et al. (2011) Anti-beta-amyloid immunotherapy for Alzheimer's disease: focus on bapineuzumab. Curr Alzheimer Res 8: 808-817. doi: 10.2174/156720511798192718
![]() |
[130] |
Lambracht-Washington D, Qu BX, Fu M, et al. (2011) DNA immunization against amyloid beta 42 has high potential as safe therapy for Alzheimer's disease as it diminishes antigen-specific Th1 and Th17 cell proliferation. Cell Mol Neurobiol 31: 867-874. doi: 10.1007/s10571-011-9680-7
![]() |
[131] |
Salloway S, Sperling R, Brashear HR (2014) Phase 3 trials of solanezumab and bapineuzumab for Alzheimer's disease. N Engl J Med 370: 1460. doi: 10.1056/NEJMc1402193
![]() |
[132] |
Salloway S, Sperling R, Fox NC, et al. (2014) Two phase 3 trials of bapineuzumab in mild-to-moderate Alzheimer's disease. N Engl J Med 370: 322-333. doi: 10.1056/NEJMoa1304839
![]() |
[133] |
Krishnamurthy PK, Sigurdsson EM (2011) Therapeutic applications of antibodies in non-infectious neurodegenerative diseases. New Biotechnol 28: 511-517. doi: 10.1016/j.nbt.2011.03.020
![]() |
[134] |
Britschgi M, Wyss-Coray T (2009) Blood protein signature for the early diagnosis of Alzheimer disease. Arch Neurol 66: 161-165. doi: 10.1001/archneurol.2008.530
![]() |
[135] |
Lindberg C, Chromek M, Ahrengart L, et al. (2005) Soluble interleukin-1 receptor type II, IL-18 and caspase-1 in mild cognitive impairment and severe Alzheimer's disease. Neurochem Int 46: 551-557. doi: 10.1016/j.neuint.2005.01.004
![]() |
[136] |
Reddy MM, Wilson R, Wilson J, et al. (2011) Identification of Candidate IgG Biomarkers for Alzheimer's Disease via Combinatorial Library Screening. Cell 144: 132-142. doi: 10.1016/j.cell.2010.11.054
![]() |
1. | Zin Thu Win, Boping Tian, Population dynamical behaviors of jellyfish model with random perturbation, 2023, 28, 1531-3492, 2994, 10.3934/dcdsb.2022201 | |
2. | Da Song, Wentao Fu, Meng Fan, Kun Li, Ecological effect of seasonally changing temperature on the life cycle of Aurelia aurita, 2025, 501, 03043800, 111014, 10.1016/j.ecolmodel.2024.111014 |
Parameter | Ranges | Ref. | Unit | data 1 | data 2 |
$ \alpha(T) $ | $ 0.03\sim0.15^{a} $ | [14] | ind $ \cdot $ d$ ^{-1} $ $ P^{-1} $ | 0.12 | 0.15 |
$ \beta(T) $ | $ 0.065\sim0.139^{a} $ | [14] | ind $ \cdot $ d$ ^{-1} $time$ ^{-1} $ $ P^{-1} $ | 0.108 | 0.122 |
$ \gamma $ | $ 19\sim178^{a} $ | [33,34] | ind $ \cdot $ d$ ^{-1} $ $ M^{-1} $ | 100 | 170 |
$ s_1 $ | $ 0.001\sim0.3^{b} $ | [34,35] | no unit | 0.008 | 0.01 |
$ s_2 $ | $ 0.01\sim0.8^{b} $ | [34] | no unit | 0.2 | 0.8 |
$ n $ | $ 1\sim2 $ | [14] | times | 1 | 1 |
$ d_1 $ | $ 0\sim0.028^{a, b} $ | [6,14] | d$ ^{-1} $ | 0.0001 | 0.0001 |
$ d_2 $ | $ 0.0001\sim0.3^{b} $ | [34] | d$ ^{-1} $ | 0.0001 | 0.0001 |
$ d_3 $ | $ 0.004\sim0.02^{a} $ | [34,36] | d$ ^{-1} $ | 0.006 | 0.004 |
$ d_4 $ | $ 0.0001\sim0.8^{b} $ | [1] | d$ ^{-1} $ | 0.0001 | 0.0001 |
$ b_1 $ | $ 0.00001\sim0.1^{b} $ | [3,37] | d$ ^{-1} $ind$ ^{-1} $ | 0.0012 | 0.0001 |
$ b_2 $ | $ 0\sim0.1^{b} $ | d$ ^{-1} $ind$ ^{-1} $ | 0.0001 | 0.0001 | |
$ \tau_1 $ | $ 30\sim120^{b} $ | [10,33] | d | 120 | 120 |
$ \tau_2 $ | $ 60\sim300^{b} $ | [10,33] | d | 90 | 150 |
Parameter | Ranges | Ref. | Unit | data 1 | data 2 |
$ \alpha(T) $ | $ 0.03\sim0.15^{a} $ | [14] | ind $ \cdot $ d$ ^{-1} $ $ P^{-1} $ | 0.12 | 0.15 |
$ \beta(T) $ | $ 0.065\sim0.139^{a} $ | [14] | ind $ \cdot $ d$ ^{-1} $time$ ^{-1} $ $ P^{-1} $ | 0.108 | 0.122 |
$ \gamma $ | $ 19\sim178^{a} $ | [33,34] | ind $ \cdot $ d$ ^{-1} $ $ M^{-1} $ | 100 | 170 |
$ s_1 $ | $ 0.001\sim0.3^{b} $ | [34,35] | no unit | 0.008 | 0.01 |
$ s_2 $ | $ 0.01\sim0.8^{b} $ | [34] | no unit | 0.2 | 0.8 |
$ n $ | $ 1\sim2 $ | [14] | times | 1 | 1 |
$ d_1 $ | $ 0\sim0.028^{a, b} $ | [6,14] | d$ ^{-1} $ | 0.0001 | 0.0001 |
$ d_2 $ | $ 0.0001\sim0.3^{b} $ | [34] | d$ ^{-1} $ | 0.0001 | 0.0001 |
$ d_3 $ | $ 0.004\sim0.02^{a} $ | [34,36] | d$ ^{-1} $ | 0.006 | 0.004 |
$ d_4 $ | $ 0.0001\sim0.8^{b} $ | [1] | d$ ^{-1} $ | 0.0001 | 0.0001 |
$ b_1 $ | $ 0.00001\sim0.1^{b} $ | [3,37] | d$ ^{-1} $ind$ ^{-1} $ | 0.0012 | 0.0001 |
$ b_2 $ | $ 0\sim0.1^{b} $ | d$ ^{-1} $ind$ ^{-1} $ | 0.0001 | 0.0001 | |
$ \tau_1 $ | $ 30\sim120^{b} $ | [10,33] | d | 120 | 120 |
$ \tau_2 $ | $ 60\sim300^{b} $ | [10,33] | d | 90 | 150 |