Dynamical behavior of networks of non-uniform Timoshenko beams system with boundary time-delay inputs

  • Received: 01 June 2010 Revised: 01 May 2011
  • Primary: 93D15, 47A75; Secondary: 35B40, 93D20.

  • The dynamical stability of planar networks of non-uniform Timoshenko beams system is considered. Suppose that the displacement and rotational angle is continuous at the common vertex of this network and the bending moment and shear force satisfies Kirchhoff's laws, respectively. Time-delay terms exist in control inputs at exterior vertices. The feedback control laws are designed to stabilize this kind of networks system. Then it is proved that the corresponding closed loop system is well-posed. Under certain conditions, the asymptotic stability of this system is shown. By a complete spectral analysis, the spectrum-determined-growth condition is proved to be satisfied for this system. Finally, the exponential stability of this system is discussed for a special case and some simulations are given to support these results.

    Citation: Zhong-Jie Han, Gen-Qi Xu. Dynamical behavior of networks of non-uniform Timoshenkobeams system with boundary time-delay inputs[J]. Networks and Heterogeneous Media, 2011, 6(2): 297-327. doi: 10.3934/nhm.2011.6.297

    Related Papers:

    [1] Kawkab Al Amri, Qamar J. A Khan, David Greenhalgh . Combined impact of fear and Allee effect in predator-prey interaction models on their growth. Mathematical Biosciences and Engineering, 2024, 21(10): 7211-7252. doi: 10.3934/mbe.2024319
    [2] Saheb Pal, Nikhil Pal, Sudip Samanta, Joydev Chattopadhyay . Fear effect in prey and hunting cooperation among predators in a Leslie-Gower model. Mathematical Biosciences and Engineering, 2019, 16(5): 5146-5179. doi: 10.3934/mbe.2019258
    [3] Dirk Stiefs, Ezio Venturino, Ulrike Feudel . Evidence of chaos in eco-epidemic models. Mathematical Biosciences and Engineering, 2009, 6(4): 855-871. doi: 10.3934/mbe.2009.6.855
    [4] Yuhong Huo, Gourav Mandal, Lakshmi Narayan Guin, Santabrata Chakravarty, Renji Han . Allee effect-driven complexity in a spatiotemporal predator-prey system with fear factor. Mathematical Biosciences and Engineering, 2023, 20(10): 18820-18860. doi: 10.3934/mbe.2023834
    [5] Yuanfu Shao . Bifurcations of a delayed predator-prey system with fear, refuge for prey and additional food for predator. Mathematical Biosciences and Engineering, 2023, 20(4): 7429-7452. doi: 10.3934/mbe.2023322
    [6] Hongqiuxue Wu, Zhong Li, Mengxin He . Dynamic analysis of a Leslie-Gower predator-prey model with the fear effect and nonlinear harvesting. Mathematical Biosciences and Engineering, 2023, 20(10): 18592-18629. doi: 10.3934/mbe.2023825
    [7] Ranjit Kumar Upadhyay, Swati Mishra . Population dynamic consequences of fearful prey in a spatiotemporal predator-prey system. Mathematical Biosciences and Engineering, 2019, 16(1): 338-372. doi: 10.3934/mbe.2019017
    [8] 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
    [9] Rongjie Yu, Hengguo Yu, Chuanjun Dai, Zengling Ma, Qi Wang, Min Zhao . Bifurcation analysis of Leslie-Gower predator-prey system with harvesting and fear effect. Mathematical Biosciences and Engineering, 2023, 20(10): 18267-18300. doi: 10.3934/mbe.2023812
    [10] Wanxiao Xu, Ping Jiang, Hongying Shu, Shanshan Tong . Modeling the fear effect in the predator-prey dynamics with an age structure in the predators. Mathematical Biosciences and Engineering, 2023, 20(7): 12625-12648. doi: 10.3934/mbe.2023562
  • The dynamical stability of planar networks of non-uniform Timoshenko beams system is considered. Suppose that the displacement and rotational angle is continuous at the common vertex of this network and the bending moment and shear force satisfies Kirchhoff's laws, respectively. Time-delay terms exist in control inputs at exterior vertices. The feedback control laws are designed to stabilize this kind of networks system. Then it is proved that the corresponding closed loop system is well-posed. Under certain conditions, the asymptotic stability of this system is shown. By a complete spectral analysis, the spectrum-determined-growth condition is proved to be satisfied for this system. Finally, the exponential stability of this system is discussed for a special case and some simulations are given to support these results.


    The effect of disease on eco-epidemiology system is a significant topic from both mathematical and ecological perspectives. The disease factor usually leads to a more complex and diverting dynamics than those in the disease-free system [1,2]. Within the interactions between predator and prey, the disease could only spread in prey or predator population, also could spread between prey and predator [3,4,5]. Birds (particularly pelicans) infect vibrio and die by preying on vibrio-infected fish (particularly tilapia) at the Salton Sea in the desert of Southern California [3], which is an example of disease spreads amongst the prey. For the disease in predator, taking fox rabies as an example, foxes (Vulpis) infect rabies and transmit to other foxes or their prey rabbits by biting in Europe and North America [6]. More relevant examples could be found in [7]. From the mathematical epidemiology point of view, one needs much more attention in the dynamics of infected predator to observe whether the presence of the prey allows the survival of a part of the predator population [8].

    A variety of diseased predator models have been proposed to study the complex interaction between prey and predator with infected diseases [2,9,10] and the reference therein. Most common epidemic model applied in predator-prey interactions is the $ SI $-type, i.e., the predator population $ Y(t) $ is divided into two sub-classes, namely susceptible predator $ S(t) $ and infected predator $ I(t) $, respectively [10,11,12]. The infection term could be mass-action term (bilinear form) $ \beta S I $ or saturation form $ \frac{\beta S I}{S+I} $ [4]. The infected predators usually behave differently with susceptible ones, and suffer an additional death rate. In a epidemic model, the global dynamics are usually determined by the basic reproduction number $ \mathcal{R}_0 $, i.e., the disease will dies out in the population when $ \mathcal{R}_0\leq 1 $, and the disease will persist in the population when $ \mathcal{R}_0 > 1 $. However, the basic reproduction number is no longer a threshold parameter determining the global dynamics in diseased predator models, on the contrary, the dynamics are relatively comprehensive and unexpected.

    Predation is the key force in a prey-predator interaction, which could affect the size of prey population by direct hunting [9,13,14,15], and elicit a variety of anti-predator responses [16,17,18]. Consequently, prey tends to alter behaviors in a certain extent, such as change of habitat, foraging activity, vigilance, physiological changes. This anti-predator behaviors accelerate the extinction, evolution and development of prey population in the long run. Under the risk of predation, prey may reduce its foraging activity in order to stay alert, leading to starvation which impacts on population growth [19,20]. Therefore, an immediately result of anti-predator behaviors is the reduction of prey growth rate, which is the cost for prey in prey defense [19,21,22,23,24,25,26].

    Consider a simple birth-death process of the prey $ X(t) $ with the cost of anti-predator behaviors [27]:

    $ \frac{{\mathrm{d}}X}{{\mathrm{d}}t} = [F(k,Y)a]X-dX, $

    where $ X, Y $ represent the density of the prey and predator, respectively. $ a $ is the birth rate of prey, $ d $ is the natural death rate of prey. $ F(k, Y) $ accounts for the cost of anti-predator defence due to fear, the parameter $ k $ reflects the level of fear which drives anti-predator behaviors of prey. The fear factor $ F(k, Y) $ has some specific assumptions under the ecological motions, for details see [20,27].

    To derive a simple diseased predator model incorporating the anti-predator defence due to fear, we adopted the following fear effect term $ F(k, Y) $:

    $ F(k, Y) = \frac{1}{1+kY} = \frac{1}{1+k(S+I)}. $

    Based on the results in [4,9,11], we can obtain the eco-epidemiological system with cost of anti-predator behaviors as following system of nonlinear differential equations:

    $ {dXdt=rX1+k(S+I)rX2KaXS1+bX,dSdt=eaXS1+bXd1SβSI,dIdt=βSId2I,
    $
    (1.1)

    where $ X, S, I $ represent the density of prey, susceptible predator and infected predator at time $ t $, respectively. $ r $ is the intrinsic growth rate of prey, $ K $ is the carrying capacity of the prey, $ a $ is the predation coefficient, $ b $ is the predators handling time of a prey, $ e $ is the biomass conversion constant, $ \beta $ is the transmissibility coefficient. $ d_1 $ and $ d_2 $ are the mortality rates of the susceptible predator and infected predator, and naturally $ d_1 < d_2 $.

    This paper consists of six sections. In the next section, we prove the positivity and boundedness of the solution of system (1.1). In Section 3, we provide the existence conditions of the equilibria of the model. We analyze the stability of equilibria and show the occurrence of Hopf bifurcation in Section 4. In Section 5, the correctness of the theoretical proof is illustrated by numerical simulation. Finally, we summarize our results with ecological interpretations in Section 6.

    In view of the ecological significance, we only consider the solutions $ (X(t), S(t), I(t)) $ of system (1.1) on

    $ {\mathbb R}_{+}^{3} = \{(X(t), S(t), I(t))\in {\mathbb R}_+^3:X(t)\geq0,S(t)\geq0,I(t)\geq0\}. $

    Theorem 2.1. Each solution of system (1.1) with initial value $ (X(0), S(0), I(0))\in {\mathbb R}^{3}_{+} $ is positive and ultimately bounded.

    Proof. Since the right-hand side of system (1.1) is completely continuous and locally Lipschitzian on $ {\mathbb R}^{3}_{+} $, the solution $ (X(t), S(t), I(t)) $ with initial condition $ (X(0), S(0), I(0))\in {\mathbb R}^{3}_{+} $ exists and is unique on $ {\mathbb R}^{3}_{+} $.

    By integrating, it follows from system (1.1) that

    $ X(t)=X(0)exp{t0(r1+k(S(τ)+I(τ))rX(τ)KaS(τ)1+bX(τ))dτ}0,S(t)=S(0)exp{t0(eaX(τ)1+bX(τ)d1βI(τ))dτ}0,I(t)=I(0)exp{t0(βS(τ)d2)dτ}0.
    $

    Hence, the solution $ (X(t), S(t), I(t)) $ of system (1.1) with the initial condition $ (X(0), S(0), I(0))\in {\mathbb R}^{3}_{+} $ remains positive.

    From the first equation of (1.1), we can obtain

    $ \frac{{\mathrm{d}}X}{{\mathrm{d}}t} = \frac{rX}{1+k(S+I)}- \frac{rX^2}{K}- \frac{aXS}{1+bX}\leq rX- \frac{rX^2}{K} = rX\left(1- \frac{X}{K}\right), $

    then

    $ \limsup\limits_{t\rightarrow \infty}X(t)\leq K. $

    Let $ N(t) = eX(t)+S(t)+I(t) $, we can get

    $ dNdt=erX1+k(S+I)erX2Kd1Sd2IerXerX2Kd1Sd2IerX(1XK)+ed1Xd1NeK(r+d1)24rd1N,
    $

    then

    $ \limsup\limits_{t\rightarrow \infty}N(t)\leq \frac{eK(r+d_1)^2}{4rd_1}. $

    This ends the proof.

    Remark 2.2. From Theorem 2.1, we know that all positive solutions of system (1.1) with initial conditions $ (X(0), S(0), I(0))\in \mathbb{R}_+^3 $ are defined in the following positive bounded invariant:

    $ \Gamma: = \left\{(X(t),S(t),I(t))\in {\mathbb R}^{3}_{+}:0\leq X(t)\leq K,0\leq eX(t)+S(t)+I(t)\leq \frac{eK(r+d_1)^2}{4rd_1}\right\}. $

    System (1.1) possesses at most three boundary equilibria:

    (i) Trivial equilibrium: $ E_0 = (0, 0, 0); $

    (ii) Axial equilibrium: $ E_1 = (K, 0, 0); $

    (iii) Planar equilibrium: $ E_2 = (X_2, S_2, 0) $ exists if $ ea-bd_1 > 0 $ and $ K > \frac{d_1}{ea-bd_1} $, where

    $ X2=d1eabd1,S2=[K(eabd1)2+rd1ke]+[K(eabd1)2rd1ke]2+4K2kre(eabd1)32Kk(eabd1)2.
    $
    (3.1)

    For epidemic models, the most critical problem is the threshold property for the extinction and persistence of the disease, which is generally governed by the basic reproduction number $ \mathcal{R}_0 $. The basic reproduction number can be interpreted as the expected number of secondary cases produced, in a completely susceptible population, by a typical infected individual during its entire period of infectiousness. Following [28], we define the basic reproduction number for the predator population in the system (1.1) by

    $ \mathcal{R}_0: = \dfrac{\beta S_2}{d_2}, $

    where $ S_2 $ is given by (3.1).

    Next, we mainly focus on the existence of positive equilibrium $ E_3 = (X_3, S_3, I_3) $ of system (1.1). The coordinates $ X_3, S_3, I_3 $ are positive solutions to the following system of equilibrium equations:

    $ {r1+k(S3+I3)rX3KaS31+bX3=0,eaX31+bX3d1βI3=0,βS3d2=0.
    $

    Thus,

    $ S_3 = \frac{d_2}{\beta},\;\;\;I_3 = {\frac {X_3(ea-bd_{{1}})-d_{{1}}}{ \left( bX_3+1 \right) \beta}}, $

    and $ X_3 $ is the positive root of (3.2) in $ \left(X_2, +\infty\right) $:

    $ Q(X)=m3X3+m2X2+m1X+m0=0,
    $
    (3.2)

    where

    $ m3:=bβr(k(eabd1)+b(kd2+β)),m2:=βr(Kb2βk(eabd1)2b(kd2+β)+bkd1),m1:=(kd2a(eabd1)b(akd22+aβd22β2r))Kβr(kd1+kd2+β),m0:=K(ad2(kd1kd2β)β2r).
    $

    If $ ea-bd_1 > 0 $ and $ r > \dfrac{ad_2(\beta+k(d_2-d_1))}{\beta^2} $, we have

    $ m_3 < 0,\;\;\;m_0 > 0. $

    By Descartes' rule of signs, system (1.1) has at least one positive equilibrium $ E_3. $

    Hence, we have the following results on the existence of the positive equilibrium. It is worthy to note that the positive equilibrium is not unique due to the impact of fear effect $ k $.

    Theorem 3.1. If $ ea-bd_1 > 0 $ and $ r > \frac{ad_2(\beta+k(d_2-d_1))}{\beta^2} $, then system (1.1) has at least one positive equilibrium $ E_3 = (X_3, S_3, I_3), $ where $ S_3 = \frac{S_2}{\mathcal{R}_0} $, $ I_3 = {\frac {X_3(ea-bd_{{1}})-d_{{1}}}{ \left(bX_3+1 \right) \beta}} $ and $ X_3 $ is the positive root of (3.2) in $ \left(X_2, +\infty\right) $.

    Regarding the local stability of trivial equilibrium $ E_0 $ and axial equilibrium $ E_1 $, we have the following results. The proof is standard, so we omit it here.

    Theorem 4.1. For system (1.1),

    (i) The trivial equilibrium $ E_0 = (0, 0, 0) $ is unstable;

    (ii) If one of the following inequalities holds:

    (ii-1) $ ea-bd_1 < 0 $;

    (ii-2) $ ea-bd_1 > 0 $ and $ K < \frac{d_1}{ea-bd_1} $,

    then the axial equilibrium $ E_1 = (K, 0, 0) $ is stable; while $ E_1 = (K, 0, 0) $ is unstable if $ ea-bd_1 > 0 $ and $ K > \frac{d_1}{ea-bd_1} $.

    Secondly, we will show the local stability of the planar equilibrium $ E_2 $ of system (1.1). For convenience, set

    $ r1:=d2(eabd1)βe,r2:=d2(eabd1)(ea+bd1)aβe2,K1:=ea+bd1b(eabd1),K2:=βerd1(eabd1)(βerd2(eabd1)),k1:=Kb(eabd1)2(Kb(eabd1)(ea+bd1))ae2r(ea+bd1),k2:=β((eabd1)(d2(eabd1)βer)K+βerd1)(Kd2(eabd1)2+βerd1)d2.
    $
    (4.1)

    Theorem 4.2. For system (1.1), assume that $ ea-bd_{{1}} > 0 $. If one of the following inequalities holds:

    (Ⅰ) $ r\leq r_1 $ and one of the following inequalities holds:

    (Ⅰ-1) $ \frac{d_1}{ea-bd_1} < K\leq K_1 $;

    (Ⅰ-2) $ K > K_1 $ and $ k > k_1 $;

    (Ⅱ) $ r_1 < r < r_2 $ and one of the following inequalities holds:

    (Ⅱ-1) $ \frac{d_1}{ea-bd_1} < K\leq K_1 $;

    (Ⅱ-2) $ K_1 < K $ and $ k > \max\{k_1, k_2\} $;

    (Ⅲ) $ r > r_2 $ and one of the following inequalities holds:

    (Ⅲ-1) $ \frac{d_1}{ea-bd_1} < K\leq K_2 $;

    (Ⅲ-2) $ K_2 < K $ and $ k > \max\{k_1, k_2\} $,

    then equilibrium $ E_2 $ is stable; otherwise, it is unstable.

    Proof. The Jacobian matrix of system (1.1) at $ E_2 $ is given by

    $ J2=(a11a12a13a210βS200βS2d2),
    $

    where

    $ a11:=X2(rK+abS2(1+bX2)2),a12:=krX2(1+kS2)2aX21+bX2,a13:=krX2(1+kS2)2,a21:=eaS2(1+bX2)2.
    $

    Hence, the characteristic equation of $ J_2 $ is given as

    $ f(λ)(λβS2+d2)=0,
    $
    (4.2)

    where

    $ f(\lambda): = \lambda^2-a_{11}\lambda-a_{12}a_{21}. $

    Clearly, one can see that $ J_2 $ has three eigenvalues $ \lambda_1 $, $ \lambda_2 $ and $ \lambda_3 = \beta S_2-d_2. $ Since $ a_{12} < 0, a_{21} > 0 $, then $ -a_{12}a_{21} > 0 $.

    From (3.1), we can obtain

    $ a11=X2(rK+abS2(1+bX2)2)=X2Φ2Kka2e2,
    $

    where

    $ \Phi: = ab\sqrt { \left( K \left( ea-bd_{{1}} \right) ^{2}-rd_{{1}}ke \right) ^{ 2}+4\,{K}^{2}kre \left( ea-bd_{{1}} \right) ^{3}} \\- ab(K(ea-bd_1)^2+rd_1ke)-2rka^2e^2. $

    Note that the sign of $ \Phi $ depends on

    $ ˜Φ:=a2b2((K(eabd1)2rd1ke)2+4K2kre(eabd1)3)(ab(K(aebd1)2+rd1ke)+2rka2e2)2=4a2kerP(k),
    $

    where

    $ P(k): = -a{e}^{2}r \left( ea+bd_{{1 }} \right)k+Kb(ea-bd_1)^2(Kb(ea-bd_1)-(ea+bd_1)) . $

    One can obtain that $ P(k) $ is decreasing with respect to $ k $. If $ K\leq{\frac {ea+bd_{{1}}}{b \left(ea-bd_{{1}} \right) }} $ holds, we have $ P(0)\leq0 $, which means that $ P(k) < 0 $ for all $ k > 0 $; if $ K > {\frac {ea+bd_{{1}}}{b \left(ea-bd_{{1}} \right) }} $ and $ k > k_1 $ hold, we can get $ P(k) < 0 $. Therefore, when one of the following inequalities holds:

    (i) $ K\leq{\frac {ea+bd_{{1}}}{b \left(ea-bd_{{1}} \right) }} $;

    (ii) $ K > {\frac {ea+bd_{{1}}}{b \left(ea-bd_{{1}} \right) }} $ and $ k > k_1 $,

    we can obtain $ a_{11} < 0 $, which implies that the real parts of $ \lambda_1 $ and $ \lambda_2 $ are all negative.

    It follows from system (3.1) that

    $ βS2d2=β[K(eabd1)2+rd1ke]+β[K(eabd1)2rd1ke]2+4K2kre(eabd1)32Kk(eabd1)2d2=Θ2Kk(eabd1)2,
    $

    where

    $ \Theta: = -K \left( ea-bd_{{1}} \right) ^{2} \left( 2\,kd_{{2}}+\beta \right) -\beta\,ekrd_{{1}}+\beta\,\sqrt { \left( K \left( ea-bd_{{1}} \right) ^{2}-rd_{{1}}ke \right) ^{2}+4\,{K}^{2}kre \left( ea-bd_{{1}} \right) ^{3}}. $

    Note that the sign of $ \Theta $ depends on

    $ ˜Θ:=β2(K(eabd1)2rd1ke)2+4β2K2kre(eabd1)3(K(eabd1)2(2kd2+β)+βekrd1)2=4Kk(aebd1)2[(Kd2(eabd1)2+βerd1)d2k+β((eabd1)(d2(eabd1)βer)K+βerd1)].
    $

    Then if one of the following inequalities holds:

    (Ⅰ) $ ea-bd_{{1}} > 0 $ and $ r\leq\frac{d_2(ea-bd_1)}{\beta e} $;

    (Ⅱ) $ ea-bd_{{1}} > 0 $, $ r > \frac{d_2(ea-bd_1)}{\beta e} $ and one of the following inequalities:

    (Ⅱ-1) $ K\leq\frac{\beta erd_1}{(ea-bd_1)(\beta er-d_2(ea-bd_1))} $;

    (Ⅱ-2) $ K > \frac{\beta erd_1}{(ea-bd_1)(\beta er-d_2(ea-bd_1))} $ and $ k > k_2: = -\frac{\beta((ea-bd_1)(d_2(ea-bd_1)-\beta er)K+\beta erd_1)}{(Kd_2(ea-bd_1)^2+\beta erd_1)d_2} $,

    we have $ \lambda_3 = \beta S_2-d_2 < 0. $

    Thus, we can arrive at the conclusion.

    It should be pointed out that another way to state Theorem 4.2 is as follows.

    Remark 4.3. For system (1.1), assume that $ ea-bd_1 > 0 $ and $ \mathcal{R}_0 < 1 $. If one of the following inequalities:

    (Ⅰ) $ \dfrac{d_1}{ea-bd_1} < K\leq K_1; $

    (Ⅱ) $ K > K_1 $ and $ k > k_1 $

    holds, then the planar equilibrium $ E_2 $ is stable; otherwise, it is unstable.

    Next, we will show the local stability of the positive equilibrium $ E_3 $ of system (1.1).

    The Jacobian matrix of system (1.1) at $ E_3 $ is given by

    $ J_3 = \left( b11b12b13b210d20βI30
    \right), $

    where

    $ b11=X3(rK+abS3(1+bX3)2),b12=krX3(1+k(S3+I3))2aX31+bX3<0,b13=krX3(1+k(S3+I3))2<0,b21=eaS3(1+bX3)2>0.
    $
    (4.3)

    The characteristic equation of $ J_3 $ is given as

    $ λ3+A1λ2+A2λ+A3=0,
    $
    (4.4)

    where

    $ A1=b11,A2=βd2I3b12b21,A3=b11βd2I3b13b21βI3.
    $
    (4.5)

    Note that if $ A_1 > 0 $ holds, then $ b_{11} < 0 $, which means that $ A_3 > 0. $ According to Routh-Hurwitz criterion, the positive equilibrium $ E_3 $ is locally asymptotically stable when $ A_1 > 0 $ and $ A_1A_2-A_3 > 0 $.

    Therefore, we can establish the following statement.

    Theorem 4.4. Assume that $ ea-bd_1 > 0 $ and $ r > \frac{ad_2(\beta+k(d_2-d_1))}{\beta^2} $ hold. The positive equilibrium $ E_3 $ of system (1.1) is locally asymptotically stable if $ A_1 > 0 $ and $ A_1A_2-A_3 > 0 $, where $ A_i, i = 1, 2, 3 $ is defined as in (4.5). Otherwise, it is unstable.

    Remark 4.5. Theorem 4.4 gives a sufficient condition about the stability of the positive equilibrium $ E_3 $ for system (1.1). However, the complexity of model (1.1) leads to the failure to theoretically demonstrate how the fear factor affects the stability of the positive equilibrium. This will be discussed later through numerical simulations.

    In this subsection, we take $ k $ as the bifurcation parameter. The characteristic equation of system (1.1) at $ E_3 $ is (4.4), and $ A_i(k), i = 1, 2, 3 $ are defined as (4.5).

    Theorem 4.6. Hopf bifurcation near the positive equilibrium $ E_3 $ for system (1.1) occurs whenever the critical parameter $ k $ attains the value $ k = k_h $ in the domain:

    $ \Omega = \left\{k_h\in \mathbb{R}^+: \Delta(k_h): = [A_1(k)A_2(k)-A_3(k)]|_{k = k_h} = 0\; with\; A_2(k_h) > 0,\; \left[\frac{{\mathrm{d}}\Delta(k)}{{\mathrm{d}}k}\right]\Bigg|_{k = k_h}\neq0\right\}. $

    Proof. If $ k = k_h $, the characteristic Eq (4.4) equals

    $ λ3+A1(kh)λ2+A2(kh)λ+A3(kh)=0,
    $
    (4.6)

    then (4.6) can be factorized as

    $ (λ2+A2(kh))(λ+A1(kh))=0.
    $
    (4.7)

    Clearly, (4.7) has three roots: $ \lambda_1 = i\sqrt{A_2(k_h)} $, $ \lambda_2 = -i\sqrt{A_2(k_h)} $ and $ \lambda_3 = -A_1(k_h) $. The roots are of the form $ \lambda_1 = p_1(k)+ip_2(k) $, $ \lambda_2 = p_1(k)-ip_2(k) $ and $ \lambda_3 = -p_3(k) $, where $ p_i(k) (i = 1, 2, 3) $ are real numbers.

    From the characteristic Eq (4.4), we can get

    $ dλdk=λ2A1+λA2+A33λ2+2A1λ+A2,
    $
    (4.8)

    where $ ' = \frac{{\mathrm{d}}}{{\mathrm{d}}k} $. Substituting $ \lambda = i\sqrt{A_2} $ into (4.8), we obtain that

    $ A3A2A1+iA2A22(A2iA1A2)=dΔ(k)dk2(A21+A2)+i[A2A22A2A1A2dΔ(k)dk2A2(A21+A2)],
    $

    which implies that

    $ \left[\dfrac{{\mathrm{d}}{\mathrm{Re}}(\lambda)}{{\mathrm{d}}k}\right]\Big|_{k = k_h} = -\dfrac{\frac{{\mathrm{d}}\Delta(k)}{{\mathrm{d}}k}}{2(A_1^2+A_2)}\Big|_{k = k_h}. $

    By using monotonicity condition in the real part of the complex root $ \frac{{\mathrm{d}} {\mathrm{Re}}(\lambda)}{{\mathrm{d}}k}|_{k = k_h}\neq0 $, the transversality condition $ \frac{{\mathrm{d}}\Delta(k)}{{\mathrm{d}}k}|_{k = k_h}\neq0 $ can be obtained to ensure the existence of Hopf bifurcation.

    Results from numerical simulations are provided in this section to demonstrate our theoretical results. As we will show, the observations shed lights on the impact of fear factor. We choose the parameters of system (1.1) as follows:

    $ r=0.8,a=0.2,b=0.1,e=0.9,d1=0.05,β=0.1,d2=0.053.
    $
    (5.1)

    Then we have

    $ eabd1=0.175>0,d1eabd1=0.286,r1=d2(eabd1)βe=0.103,r2=d2(eabd1)(ea+bd1)aβe2=0.106,K1=ea+bd1b(eabd1)=10.571,K2=βerd1(eabd1)(βerd2(eabd1))=0.328.
    $

    Example 5.1 (The stability of $ E_1 $).

    We adopt $ K = 0.2, k = 0.01 $, then system (1.1) has trivial equilibrium $ E_0 = (0, 0, 0) $ and axial equilibrium $ E_1 = (0.2, 0, 0) $. In this case, one can know that the conditions of Theorem 4.1 are satisfied, which means that $ E_1 $ is locally asymptotically stable. The numerical results are shown in Figure 1.

    Figure 1.  Population dynamics of $ X(t) $, $ S(t) $ and $ I(t) $ of system (1.1) with $ K = 0.2, k = 0.01 $.

    Example 5.2 (The impacts of $ K $ and $ k $ on the stability of $ E_2 $).

    In this example, we will choose three values of carrying capacity $ K $ for numerical experiments. We conclude that the carrying capacity and fear effect are other key factors related to the extinction of infected predators, in addition to the basic reproduction number $ \mathcal{R}_0 $.

    Firstly, we take $ K = 0.3 < K_2 $, then we have $ k = 0.1 $ which yields that $ \mathcal{R}_0 = 0.263 < 1 $. In this case, system (1.1) has trivial equilibrium $ E_0 = (0, 0, 0) $, axial equilibrium $ E_1 = (0.3, 0, 0) $, and planar equilibrium $ E_2 = (0.286, 0.139, 0) $. By Theorem 4.5, $ E_2 $ is locally asymptotically stable, see Figure 2(a). Thus, when the carrying capacity of the prey $ K $ is small, no matter what the level of fear $ k $ is, the small size of prey population will lead to the extinction of infected predators.

    Figure 2.  The impacts of $ K $ and $ k $ on the stability of $ E_2 $.

    Secondly, for comparison, we take $ K_2 < K = 15 $, then

    $ k1=Kb(eabd1)2(Kb(eabd1)(ea+bd1))ae2r(ea+bd1)=0.149,k2=β((eabd1)(d2(eabd1)βer)K+βerd1)(Kd2(eabd1)2+βerd1)d2=10.873.
    $

    Choosing $ k = 0.1 < \max\{k_1, k_2\} $ which yields $ \mathcal{R}_0 = 5.792 > 1 $, then we have

    $ A_1 = 0.25791 > 0,\;A_1A_2-A_3 = 0.00523 > 0. $

    In this case, system (1.1) has trivial equilibrium $ E_0 = (0, 0, 0) $, axial equilibrium $ E_1 = (15, 0, 0) $, planar equilibrium $ E_2 = (0.286, 3.070, 0) $, and positive equilibrium $ E_3 = (7.351, 0.530, 7.125) $. By Theorem 4.5, $ E_2 = (0.286, 3.070, 0) $ is unstable. On the contrary, $ E_3 = (7.351, 0.530, 7.125) $ is locally asymptotically stable. The numerical simulation is shown in Figure 2(b).

    Finally, we take $ K_2 < K = 60 $, then we have

    $ k1=Kb(eabd1)2(Kb(eabd1)(ea+bd1))ae2r(ea+bd1)=6.629,k2=β((eabd1)(d2(eabd1)βer)K+βerd1)(Kd2(eabd1)2+βerd1)d2=12.238.
    $

    Choosing $ k = 30 > \max\{k_1, k_2\} $ which yields that $ \mathcal{R}_0 = 0.649 < 1 $, system (1.1) has trivial equilibrium $ E_0 = (0, 0, 0) $, axial equilibrium $ E_1 = (60, 0, 0) $, and planar equilibrium $ E_2 = (0.286, 0.344, 0) $. By Theorem 4.5, $ E_2 $ is locally asymptotically stable, see Figure 2(c). Thus, when the carrying capacity of the prey $ K $ is relatively large, a high level of fear $ k $ will lead to the extinction of infected predators.

    Example 5.3 (The impact of $ k $ on the stability of $ E_3 $). We adopt $ K = 60 $, then we have $ k_h = 0.26 $. In the next, we will choose three values of $ k $, corresponding to the local stability of $ E_3 $, Hopf bifurcation, and instability of $ E_3 $, to illustrate the impact of fear factor on the population dynamics.

    Firstly, we take $ k = 0.1 < k_h $ which yields that $ \mathcal{R}_0 = 5.881 > 1 $, then system (1.1) has trivial equilibrium $ E_0 = (0, 0, 0) $, axial equilibrium $ E_1 = (60, 0, 0) $, planar equilibrium $ E_2 = (0.286, 3.117, 0) $ and a unique positive equilibrium $ E_3 = (24.047, 0.530, 12.213) $. In this case, we obtain that

    $ A_1 = 0.29863 > 0,\;A_1A_2-A_3 = 0.00065 > 0, $

    which means that $ E_3 $ is local asymptotically stable. The numerical results are shown in Figure 3.

    Figure 3.  Population dynamics of $ X(t) $, $ S(t) $ and $ I(t) $ of system (1.1) with $ K = 60, k = 0.1 < \; k_h $.

    Secondly, we take $ k = 0.26 = k_h $ which yields that $ \mathcal{R}_0 = 4.682 > 1 $, then system (1.1) has trivial equilibrium $ E_0 = (0, 0, 0) $, axial equilibrium $ E_1 = (60, 0, 0) $, planar equilibrium $ E_2 = (0.286, 2.481, 0) $ and a unique positive equilibrium $ E_3 = (12.975, 0.530, 9.665) $. In this case, we obtain that

    $ A_1 = 0.14694 > 0,\;A_1A_2-A_3 = 0, $

    which means that system (1.1) undergoes a Hopf bifurcation and there is a limit cycle around $ E_3 $. The numerical results and the bifurcation diagrams of system (1.1) with respect to the parameter $ k $ are shown in Figures 4 and 5, respectively. Comparing Figures 3 and 4(a), one can see that there are two different implications induced by the fear factor $ k $: the first is that the stability of $ E_3 $ converts from stable into unstable, and the second is the decrease of values of $ X_3 $ and $ I_3 $ of $ E_3 $.

    Figure 4.  Population dynamics of $ X(t) $, $ S(t) $ and $ I(t) $ of system (1.1) with $ K = 60, k = 0.26 = k_h $. (a) Time-series plots; (b) Phase portraits in 3-dimensional space.
    Figure 5.  Bifurcation diagram of the system (1.1) with respect to the parameter $ k $. Here $ K = 60 $, other parameters are taken as in (5.1).

    Finally, we take $ k = 0.5 > k_h $ which yields that $ \mathcal{R}_0 = 3.821 > 1 $, then system (1.1) has trivial equilibrium $ E_0 = (0, 0, 0) $, axial equilibrium $ E_1 = (60, 0, 0) $, planar equilibrium $ E_2 = (0.286, 2.025, 0) $ and a unique positive equilibrium $ E_3 = (7.685, 0.530, 7.322) $. In this case, we can obtain that

    $ A_1 = 0.07642 > 0,\;A_1A_2-A_3 = -0.00051 < 0, $

    which means that $ E_3 $ is unstable. The numerical results are shown in Figure 6. One can find that the difference between Figures 4 and 6 is the decrease of values of $ E_3 $ from $ (12.975, 0.530, 9.665) $ to $ (7.685, 0.530, 7.322), $ which is induced by the impact of the feat factor.

    Figure 6.  Population dynamics of $ X(t) $, $ S(t) $ and $ I(t) $ of system (1.1) with $ K = 60, k = 0.5 > k_h $. (a) Time-series plots; (b) Phase portraits in 3-dimensional space.

    In this paper, we explored a predator-prey model that incorporates infectious disease in predator population and the cost of anti-predator behaviors. The cost of anti-predator behaviors is measured by a fear effect $ k $ leading to an reduction of prey's birth rate. We fulfill a complete stability analysis of equilibria for system (1.1) and show that the system (1.1) exhibits the Hopf bifurcation. Biologically, we focus on the impact of fear effect on the population dynamics. As we will see later, the cost of a high level of fear effect is disastrously. The main findings are summarized in the following.

    $ 1) $ Small size of prey population leads to the extinction of infected predators.

    If the carrying capacity $ K $ is relatively small, the planar equilibrium $ E_2 $ is stable, see Figure 2(a). Thus, no matter what the level of fear effect $ k $ is, a small size of prey population will lead to the extinction of infected predators.

    $ 2) $ Low level of the fear effect doesn't impact on the population dynamics.

    If the level of fear effect $ k < k_h $, the positive (coexistence) equilibrium $ E_3 $ is stable, see Figure 3. Hence, we conclude that a small fear effect $ k $ is not the key disturbance and does not change the coexistence dynamics of system (1.1). However, the densities of the prey and infected predator gradually decrease as $ k $ increasing.

    $ 3) $ Certain medium level of the fear effect lead to periodic oscillation.

    If $ k = k_h $, the fear effect can destabilize the stability of $ E_3 $ and will benefit the occurrence of periodic oscillation. In other words, system (1.1)undergoes a limit cycle, see Figures 4 and 5.

    $ 4) $ High level of the fear effect leads to complex dynamics and the infected predator can go to extinction.

    If $ k > k_h $, $ E_3 $ is unstable, see Figure 6. Therefore, a large fear effect $ k $ persistently and dramatically influence the population dynamics of prey and predator. Furthermore, if the level of the fear factor $ k $ is extremely high, the planar equilibrium $ E_2 $ is stable, see Figure 2(c). The prey will respond to perceived predation risk and show a variety of anti-predator responses, dramatically decreasing the recruitment of susceptible predator, which will lead to an extinction of infected predator.

    The authors would like to thank the editor and the referees for their helpful comments. This research was supported by the National Natural Science Foundation of China (Grant No. 12171192, 12031020 and 12071173), the Science and Technology Research Projects of the Education Office of Jilin Province, China (JJKH20211033KJ), the Technology Development Program of Jilin Province, China (20210508024RQ) and Huaian Key Laboratory for Infectious Diseases Control and Prevention, China (HAP201704).

    The authors declare that there are no conflicts of interest regarding the publication of this paper.

    [1] R. A. Adams, "Sobolev Spaces," Pure and Applied Mathematics, Vol. 65, Academic Press, New York, 1975.
    [2] K. Ammari, Asymptotic behaviour of some elastic planar networks of Bernoulli-Euler beams, Appl. Anal., 86 (2007), 1529-1548. doi: 10.1080/00036810701734113
    [3] K. Ammari and M. Jellouli, Stabilization of star-shaped networks of strings, Differential and Integral Equations, 17 (2004), 1395-1410.
    [4] K. Ammari and M. Jellouli, Remark on stabilization of tree-shaped networks of strings, Applications of Mathematics, 52 (2007), 327-343. doi: 10.1007/s10492-007-0018-1
    [5] K. Ammari, M. Jellouli and M. Khenissi, Stabilization of generic trees of strings, Journal of Dynamical and Control Systems, 11 (2005), 177-193. doi: 10.1007/s10883-005-4169-7
    [6] S. A. Avdonin and S. A. Ivanov, "Families of Exponentials. The Method of Moments in Controllability Problems for Distributed Parameter Systems," Cambridge University Press, Cambridge, 1995.
    [7] J. W. Brown and R. V. Churchill, "Complex Variables and Applications," Seventh Edition, China Machine Press, Beijing, 2004.
    [8] P. G. Casazza and G. Kutyniok, Frames of subspaces, Contemp. Math., 345 (2004), 87-113.
    [9] G. Chen, M. Coleman and H. H. West, Pointwise stabilization in the middle of the span for second order systems, nonuniform and uniform exponential decay of solutions, SIAM J. Appl. Math., 47 (1987), 751-780. doi: 10.1137/0147052
    [10] G. Chen, M. Delfour, A. Krall and G. Payre, Modeling, stabilization and control of seraially connected beams, SIAM J. Control Optim, 25 (1987), 526-546. doi: 10.1137/0325029
    [11] G. Chen, S. G. Krantz, D. L. Russell, C. E. Wayne, H. H. West and M. P. Coleman, Analysis, designs, and behavior of dissipative joints for coupled beams, SIAM J. Appl. Math., 49 (1989), 1665-1693. doi: 10.1137/0149101
    [12] R. Datko, Two examples of ill-posedness with respect to small time delays in stabilized elastic systems, IEEE Trans. Automatic Control, 38 (1993), 163-166. doi: 10.1109/9.186332
    [13] R. Datko, J. Lagnese and M. P. Polis, An example on the effect of time delays in boundary feedback stabilization of wave equations, SIAM J. Control Optim., 24 (1986), 152-156. doi: 10.1137/0324007
    [14] I. C. Gohberg and M. G. Krein, "Introduction to the Theory of Linear Nonselfadjoint Operators," AMS Transl. Math. Monographs, American Mathematical Society, 1969.
    [15] B. Z. Guo and Y. Xie, A sufficient condition on Riesz basis with parentheses of non-self-adjoint operator and application to a serially connected string system under joint feedbacks, SIAM J. Control Optim., 43 (2004), 1234-1252. doi: 10.1137/S0363012902420352
    [16] B. Z. Guo and K. Y. Yang, Output feedback stabilization of a one-dimensional Schrödinger equation by boundary observation with time delay, IEEE Transactions on Automatic Control, 55 (2010), 1226-1232. doi: 10.1109/TAC.2010.2051070
    [17] Z. J. Han and L. Wang, Riesz basis property and stability of planar networks of controlled strings, Acta Appl. Math., 110 (2010), 511-533. doi: 10.1007/s10440-009-9459-8
    [18] Z. J. Han and G. Q. Xu, Spectrum and dynamical behavior of a kind of planar network of non-uniform strings with non-collocated feedbacks, Networks and Heterogeneous Media, 5 (2010), 315-334. doi: 10.3934/nhm.2010.5.315
    [19] Z. J. Han and G. Q. Xu, Exponential stabilisation of a simple tree-shaped network of Timoshenko beams system, International Journal of Control, 83 (2010), 1485-1503. doi: 10.1080/00207179.2010.481767
    [20] Z. J. Han, G. Q. Xu, Stabilization and Riesz basis of a star-shaped network of Timoshenko beams, Journal of Dynamical and Control Systems, 16 (2010), 227-258. doi: 10.1007/s10883-010-9091-y
    [21] Z. J. Han and G. Q. Xu, Stabilization and Riesz basis property of two serially connected Timoshenko beams system, Z. Angew. Math. Mech., 89 (2009), 962-980. doi: 10.1002/zamm.200800176
    [22] Z. J. Han and G. Q. Xu, Exponential stability of Timoshenko beam system with delay terms in boundary feedbacks, ESAIM: Control, Optimisation and Calculus of Variations, 17 (2011), 552-574. doi: 10.1051/cocv/2010009
    [23] J. Lagnese, G. Leugering and E. J. P. G. Schmidt, "Modeling, Analysis of Dynamic Elastic Multi-Link Structures," Birkhäuser-Verlag, Boston-Basel-Berlin, 1994.
    [24] J. S. Liang and Y. Q. Chen, Boundary control of wave equations with delayed boundary measurement, Proceedings of IEEE International Conference on Robotics and Biomimetics, 2004, Shenyang, China, 849-854. doi: 10.1109/ROBIO.2004.1521895
    [25] J. S. Liang, Y. Q. Chen and B. Z. Guo, A new boundary control method for beam equation with delayed boundary measurement using modified smith predictors, Proceedings of the 42nd IEEE Conference on Decision and Control, 2003, Hawaii, USA, 809-814.
    [26] Yu. I. Lyubich and V. Q. Phóng, Asymptotic stability of linear differential equations in Banach spaces, Studia Math., 88 (1988), 34-37.
    [27] R. Mennicken and M. Möller, "Non-self-adjoint Boundary Eigenvalue Problem," North-Holland Mathematics Studies, vol. 192, North-Holland Publishing Co., Amsterdam, 2003.
    [28] D. Mercier, Spectrum analysis of a serially connected Euler-Bernoulli beams problems, Networks and Heterogeneous Media, 4 (2009), 709-730. doi: 10.3934/nhm.2009.4.709
    [29] D. Mercier and V. Régnier, Spectrum of a network of Euler-Bernoulli beams, Journal of Mathematical Analysis and Applications, 337 (2008), 174-196. doi: 10.1016/j.jmaa.2007.03.080
    [30] D. Mercier and V. Régnier, Control of a network of Euler-Bernoulli beams, Journal of Mathematical Analysis and Applications, 342 (2008), 874-894. doi: 10.1016/j.jmaa.2007.12.062
    [31] W. Michiels and S. I. Niculescu, "Stability and Stabilization of Time-Delay Systems. An Eigenvalue-Based Approach," Society for Industrial and Applied Mathematics, Philadelphia, 2007. doi: 10.1137/1.9780898718645
    [32] O. Morgul, On the stabilization and stability robustness against small delays of some damped wave equation, IEEE Trans. Automatic Control, 40 (1995), 1626-1630. doi: 10.1109/9.412634
    [33] S. Nicaise and C. Pignotti, Stability and instability results of the wave equation with a delay term in the boundary or internal feedbacks, SIAM J. Control Optim., 45 (2006), 1561-1585. doi: 10.1137/060648891
    [34] S. Nicaise and C. Pignotti, Stabilization of the wave equation with boundary or internal distributed delay, Differential and Integral Equations, 21 (2008), 935-958.
    [35] S. Nicaise and J. Valein, Stabilization of the wave equation on 1-D networks with a delay term in the nodal feedbacks, Networks and Heterogeneous Media, 2 (2007), 425-479. doi: 10.3934/nhm.2007.2.425
    [36] A. Pazy, "Semigroups of Linear Operators and Applications to Partial Differential Equations," Springer-Verlag, Berlin, 1983.
    [37] A. A. Shkalikov, Boundary problems for ordinary differential equations with parameter in the boundary conditions, J. Soviet Math., 33 (1986), 1311-1342. doi: 10.1007/BF01084754
    [38] K. Sriram and M. S. Gopinathan, A two variable delay model for the circadian rhythm of Neurospora crassa, J. Theor. Biol., 231 (2004), 23-38. doi: 10.1016/j.jtbi.2004.04.006
    [39] J. Srividhya and M. S. Gopinathan, A simple time delay model for eukaryotic cell cycle, Journal of Theoretical Biology, 241 (2006), 617-627. doi: 10.1016/j.jtbi.2005.12.020
    [40] H. Suh and Z. Bien, Use of time-delay actions in the controller design, IEEE Trans. Automatic Control, 25 (1980), 600-603. doi: 10.1109/TAC.1980.1102347
    [41] S. Timoshenko, "Vibration Problems in Engineering," Van Norstrand, New York, 1955.
    [42] J. Valein and E. Zuazua, Stabilization of the wave equation on 1-d networks, SIAM J. Contr. Optim, 48 (2009), 2771-2797. doi: 10.1137/080733590
    [43] Q. P. Vu, J. M. Wang, G. Q. Xu and S. P. Yung, Spectral analysis and system of fundamental solutions for Timoshenko beams, Appl. Math. Lett., 18 (2005), 127-134. doi: 10.1016/j.aml.2004.09.001
    [44] J. M. Wang and B. Z. Guo, Riesz basis and stabilization for the flexible structure of a symmetric tree-shaped beam network, Math. Meth. Appl. Sci., 31 (2008), 289-314. doi: 10.1002/mma.909
    [45] G. Q. Xu, B. Z. Guo, Riesz basis property of evolution equations in Hilbert spaces and application to a coupled string equation, SIAM J. Control Optim., 42 (2003), 966-984. doi: 10.1137/S0363012901400081
    [46] G. Q. Xu, Z. J. Han and S. P. Yung, Riesz basis property of serially connected Timoshenko beams, International Journal of Control, 80 (2007), 470-485. doi: 10.1080/00207170601100904
    [47] G. Q. Xu and J. G. Jia, The group and Riesz basis properties of string systems with time delay and exact controllability with boundary control, IMA Journal of Mathematical Control and Information, 23 (2006), 85-96.
    [48] G. Q. Xu, D. Y. Liu and Y. Q. Liu, Abstract second order hyperbolic system and applications to controlled networks of strings, SIAM J. Control Optim., 47 (2008), 1762-1784. doi: 10.1137/060649367
    [49] G. Q. Xu and S. P. Yung, The expansion of semigroup and criterion of Riesz basis, Journal of Differential Equations, 210 (2005), 1-24. doi: 10.1016/j.jde.2004.09.015
    [50] G. Q. Xu, S. P. Yung and L. K. Li, Stabilization of wave systems with input delay in the boundary control, ESAIM: Control, Optimisation and Calculus of Variations, 12 (2006), 770-785. doi: 10.1051/cocv:2006021
    [51] R. M. Young, "An Introduction to Nonharmonic Fourier Series," Pure and Applied Mathematics, vol. 93, Academic Press, London, 1980.
  • This article has been cited by:

    1. Chunmei Zhang, The effect of the fear factor on the dynamics of an eco-epidemiological system with standard incidence rate, 2024, 9, 24680427, 128, 10.1016/j.idm.2023.12.002
    2. Hongqiuxue Wu, Zhong Li, Mengxin He, Dynamic analysis of a Leslie-Gower predator-prey model with the fear effect and nonlinear harvesting, 2023, 20, 1551-0018, 18592, 10.3934/mbe.2023825
    3. Zhuoying Zhao, Xinhong Zhang, Unraveling the transmission mechanism of animal disease: Insight from a stochastic eco-epidemiological model driven by Lévy jumps, 2025, 191, 09600779, 115859, 10.1016/j.chaos.2024.115859
    4. Huazhou Mo, Yuanfu Shao, Stability and bifurcation analysis of a delayed stage-structured predator–prey model with fear, additional food, and cooperative behavior in both species, 2025, 2025, 2731-4235, 10.1186/s13662-025-03879-y
  • Reader Comments
  • © 2011 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(3979) PDF downloads(90) Cited by(8)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog